You are on page 1of 259
iB. USN ues DU BETON E Tones BOON ee: Be Nod Sel) ae ase ya aoe pe 8 STATE OF THE ART COMITE EURO-INTERNATIONAL DU BETON FASTENINGS TO CONCRETE _ AND MASONRY _ STRUCTURES STATE OF THE ART REPORT | : — : : 4 Thomas Telford Published by Thomas Telford Services Ltd, Thomas Telford House, 1 Heron Quay, London E14 45D, UK, for the Comité Euro-International du Béton, Case Postale 88, CH-1015 Lausanne, Switzerland First published 1994 Distributors for Thomas Telford books are USA: American Society of Civil Engineers, Publications Sales Department, 345 Bast 47th Street, New York, NY 10017-2398 Japan: Maruzen Co. Lid, Book Department, 3-10 Nihonbashi 2-chome, ‘Chuo-ku, Tokyo 103 Australia: DA Books and Journals, 648 Whitehorse Road, Mitcham 3132, Victoria A catalogue record fortis book is aallable from the Bish Library Casicaton ‘Avalablliy: Uncesticted ‘Content: Guidance based on research and best curren practice ‘Stott: Commitee guided {User Sructucl engnons, signers ISBN: 0 7277 1937 8 though the Comité Buo-Inerstional du Bétn and Thomas Telford Services Ld have done thir best, {o ensue tht any information given is accurate, no laity or responsibility of any kind canbe arceptet inthis respect by the Conte, Thomas Telford, their members, their servents or their agents © Comité Euto-ntermaionl du Béton, 1991 © tis presentation Thomas Telford Sorvices Lu, 1994 AT rights, including translation ceserved. Except for fir copying, no pat ofthis publication may be ‘eprcdiced,sored ne etival system or canst in ny form orb any rears, electronic, mechanical, ‘photocopying or otherwise, ido the price writen persion ofthe Publications Manage, Publications Divison, Thomas Telford Services Lil, Thomas Teforé House, | Heroa Quay, London B16 40D, “Typeset in Great Belusin by MHL Typeseting Limite, Coventry Drined and bound in Gret Britain by Redwood Press Lid, Melistam, Wilts oes centre Preface Modern fastening technique is increasingly applied for the transfer of concentrated loads into concrete and masonry structures. Cast-in-place systems (which are placed in the formwork before casting of the concrete) and post- installed systems (which are installed in hardened structural concrete or masonry) are used. ‘The ioad is transferred into the base material by mechanical interlock, friction, bond or a combination of these mechanisms. In general independent of the load-transfér mechanism, fastenings rely on the concrete tensile capacity Although a large number of fastening assemblies are installed every day, knowledge in the engineering profession about their behaviour is generally very limited. Furthermore, there is no generally accepted design method. Tn order to improve the situation, Task Group VI/S: Fastenings to Reinforced Concrete and Masonry Structures has been formed by CEB (@ to compile and compare the available research results on the behaviour of fastening systems () to propose a consistent approach based on current empirical and theoretical models for the design of fastenings (6) to develop design methods that account for the effects of fastenings and the loads they carry on the behaviour of the structures to which they are attached. ‘The Task Group has had meetings in Stuttgart (November 1987) Orlando (March 1988) Dubrovnik (September 1988) Berlin (May 1989) Toronto (March 1990) Paris (September 1990) Tokyo (May 1991). In this report, the behaviour of fastenings in concrete and masonry for the entire range of loading types (including monotonic, sustained, fatigue, seismic and impact loading), as well as the influence of environmental effects, is reviewed based on experimental results from various parts of the world. Existing theoretical approaches to prediction of the behaviour of anchors are described. The report was approved by the Task Group at its Tokyo meeting. Ina second report, the existing design approaches will be compiled and ‘compared with the available experimental data, and a consistent design method will be recommended. This work is intended to be applicable to both new constructions and the repair and strengthening of existing structures. It is not within the scope of the Task Group to review the various test methods for assessment of proper functioning (suitability) or allowable conditions of use for fastening systems. These topics are properly addressed by the appropriate regulatory bodies, for example UEAte, OTA and ASTM. ‘A report of this kind cannot be complete. Nevertheless, it is hoped that it will help the profession to understand better the behaviour of fastenings and to improve their design. Rolf Bligehausen Chairman of the Task Group Stuttgart, March 1994 Acknowledgements This report has been written by the members and associate members of the CEB Task Group VI/5: Fastenings to Reinforced Concrete and Masonry Structures 5 R. Eligehausen (Chairman), Germany R, Tewes (Technical Secretary until September 1989), Switzerland K. Bergmeister (Technical Secretary since January 1990), Italy M. Combette, France V. Covert, USA, L. Elfgren, Sweden P. Hollenbach, USA B. Kato, Japan R.E. Klingner, USA H.B. Lancelot, USA K, Laternser, Germany J. Lugez, France T. Okada, Japan W.S. Paterson, Great Britain M. Rinklake, Germany H_D. Seghezzi, Liechtenstein G. Senkiw, Canada JF, Silva, USA G. Sdderlind, Sweden R. Tewes, Switzerland J. Tschositsch, Germany E, Vintzéleou, Greece H. Wiewel, USA R.E. Wollmershauser, USA . Akiyama (Associate Member), Japan Hosokawa (Associate Member), Japan |. Kimura (Associate Member), Japan Kemer (Associate Member), Germany Matsuzaki (Associate Member), Japan Tanaka (Associate Member), Japan . Usami (Associate Member), Japan Editorial Board R, Bligehausen (Chairman), Germany K. Bergmeister (Technical Secretary), Italy R.B. Klinger, USA W.S. Paterson, Great Britain H.D, Seghezzi, Liechtenstein I.E. Silva, USA ‘The final draft has been produced in Stuttgart mainly by Konrad Bergmeister, Rolf Eligehausen and John F. Silva. Most of the figures have been drawn by Mrs Boese. Acknowledgement is due to all those who have made contributions. vRPOR< 3.3. Masonry anchors, 29 PART Il, BEHAVIOUR OF FASTENING SYSTEMS UNDER MONOTONIC LOADING IN CONCRETE — EXPERIMENTAL RESULTS 4. Introduction to fastening systems under monotonic loading in concrete scat ieneanRS: 4. 42. 43. Scope, 32 Units, 33 Cracked concrete, 33 . Headed anchors in uncracked concrete 5.1. 5.2. 53 5.4, 5.35. 5.6. Prestressing and relaxation, 37 Tension loading, 39 Shear loading, 61 Combined tension and shear loading, 75 Bending of the baseplate, 77 Behaviour of multiple-anchor ductile attachments to conerete, 78 6. Headed anchors in cracked concrete 6.1. 62. 63. Prestressing and relaxation, 84 Tension loading, 84 Shear loading, 89 7. Channel bars Tt 72. Uneracked concrete, 90 Cracked concrete, 91 8. Expansion anchors in uncracked concrete 8.1 8.2 83, 84, 85. Prestressing and relaxation, 92 Tension loading, 94 Shear loading, 99 Combined tension and shear loading, 100 Bending moment, 101 9. Expansion anchors in cracked concrete OL 9.2. 9.3. 9.4, Prestressing and relaxation, 104 Tension loading, 104 Shear loading, 110 Combined tension and shear loading, 110 10 32 37 90 92 104 CONTENTS 40. Undercut anchors in uncracked concrete 10.1. Prestressing and relaxation, 111 10.2. Tension loading, 112 10.3. Shear loading, 113 10.4. Combined tension and shear loading, 113 11. Undercut anchors in cracked concrete 1.1, Prestressing and relaxation, 114 11.2. Tension loading, 114 11.3. Shear loading, 116 11.4. Combined tension and shear loading, 116 12, Bonded anchors in uncracked concrete 12.1. Prestressing and relaxation, 117 12.2. Tension loading, 118 12,3. Shear loading, 123 12.4. Combined tension and shear loading, 124 13. Bonded anchors in cracked concrete 13.1. Prestressing and relaxation, 125 13.2. Tension loading, 125 13.3. Shear loading, 126 13.4. Combined tension and shear loading, 127 14, Plastic fasteners 14.1. Uneracked concrete, 128 14.2. Cracked concrete, 135 15. Powder-actuated fasteners 15.1. Uneracked concrete, 137 15.2. Cracked concrete, 141 PART Ill, BEHAVIOUR OF FASTENING SYSTEMS UNDER MONOTONIC LOADING IN MASONRY — EXPERIMENTAL RESULTS 16. 17. 18, 19. 20. Essential characteristics of masonry 16.1. Masonry construction in the USA, 143 16.2. Masonry construction in Europe, 146 16.3. Comparison of fastening in reinforced concrete and masonry, 148 Cast-in-place anchors 17.1, Failure loads, 149 Expansion anchors Injection-type bonded anchors in masonry 19.1. Prestressing and relaxation, 151 19.2. Tension loading, 151 19.3. Shear loading, 152 19.4. Combined tension and shear loading, 152 Plastic anchors 20.1. Prestressing and relaxation, 153 20.2. Tension loading, 153 20.3. Shear loading, 155 m1 414 47 125 128 137 143 149 150 151 153 CONTENTS PART IV. BEHAVIOUR OF FASTENING SYSTEMS UNDER SUSTAINED, CYCLIC, SEISMIC AND IMPACT LOADING 21. Types of loading and force 156 22. Sustained loading 159 22.1. Uncracked concrete, 159 22.2. Cracked concrete, 159 23. Cyclic loading 161 23.1. Introduction, 161 23.2. Prestressing, 161 23.3. Uneracked concrete, 163 23.4. Cracked concrete, 166 24, Seismic loading 167 24.1. Introduction, 167 24.2. Fastenings subjected to cyclic axial actions, 167 24.3. Fastenings subjected to cyclic shear actions, 168 24.4. Fastenings under combined tension and shear actions, 178 25. Impact loading 179 25.1. Introduction, 179 25.2. Fastenings subjected to impact tension loading, 179 25.3. Fastenings subjected to impact shear loading, 182 PART V. ENVIRONMENTAL INFLUENCES 26. Performance requirements of fasteners 183 27. Corrosion 184 27.1 Introduction, 184 27.2. Cast-in-place concreted or mortared-in components, 184 27.3. Galvanized (electroplated) components, 185 27.4, Hot-ip galvanized components, 185 27.5. Contact of galvanized components with concrete, 186 27.6. Stainless steel, 186 28. Other environmental influences 188 28.1. Humidity, 188 28.2, Temperature, 189 28.3. Ultraviolet exposure, 192 28.4, Freeze—thaw cycles, 193 28.5. Salt (corrosion) exposure, 194 28.6. Acid rain wetting and drying exposure, 194 28.7. Combination exposure, 195 29. Fire exposure 196 29.1. Introduction, 196 29.2. Experimental investigations, 198 CONTENTS PART VI. INFLUENCE OF FASTENINGS ON THE BEHAVIOUR OF THE MEMBER SERVING AS BASE MATERIAL 30. Influence of fastenings on reinforced concrete and masonry structures PART Vil. THEORETICAL MODELLING 31. Introduction to theoretical modelling 32, Behaviour of concrete under tension loading 32.1. Introduction, 205 32.2. Behaviour of concrete in tension, 206 33. Experimental investigations of the fracture of fastenings 34. Theoretical investigations 34.1. Theory of elasticity, 219 34.2. Theory of plasticity, 219 34.3. Strength criteria combined with a ‘smeared crack’ approach, 220 34.4. Fracture mechanics, 223 35. Conclusions for theoretical modelling 36. Use of the concrete tensi capacity 37. Summary References Bibliography 199 204 205 ant 219 233 234 235 236 249 1. Scope ‘The question of how best to join the many parts of a building structure is as old as building construction itself. Close examination of treditional stone, masonry and timber construction methods reveals the various and often ingenious solutions developed to transfer load from one part of a structure to another. Modern building structures, whether made of wood, masonry, concrete or steel, likewise comprise interconnected parts. The connections take many forms. In the design of concrete and masonry structures, attachments for the introduction of concentrated loads as well as primary structural connections for prefabricated components involve the use of fastening systems, These systems are also increasingly employed in the repair and strengthening of existing structures. The term ‘fastening system’ as used here encompasses the numerous methods and products currently available to facilitate the introduction of concentrated loads into concrete and masonry structures. The demands of the building industry for more flexible and dependable construction ‘methods, and efforts to improve the performance of building structures, have in recent years focused particular attention on the design, installation and performance of fastening systems. Fastening systems can be classified into cast-in-place systems and post- installed systems. Cast-in-place systems are placed in the formwork before casting of the concrete. The element to be fastened is connected to the embedded part either by the use of mechanical fasteners or by welding. Forces acting on cast-in- place systems are often distributed into the concrete member by means of suitably placed reinforcement. Post-installed systems may be installed in masonry or hardened concrete. Advances in drilling technology have led to the development of increasingly dependable post-installed systems which are widely used in new construction as well as for repair and strengthening work Each fastening system is by design suited for particular applications. The selection of the appropriate system is based on engineering, economic and architectural considerations. Modern fastening technology offers architects and engineers new versatility in that it provides systems for an increasingly wide range of uses and load levels. In order to choose the correct system for a given application and to desiga the fastening properly , comprehensive knowledge of the broad range of fastening systems and their behaviour is required. However, because the design of a particular fastening may be critical for the behaviour of the structure in which itis used, it should be based on sound engineering models that describe the real behaviour under load as closely as possible. The present state of the art cannot quantitatively account for all factors that influence the load-bearing behaviour of fastening systems. Theoretical models up to now have been based primarily on empirical test data. These limitations notwithstanding, this report compiles and compares the available experimental and theoretical research results concerning fastening systems. Ina further report, a consistent approach based on current empirical and theoretical models for the design of fastenings will be proposed, and the effects of fastenings and the loads they carry on the behaviour of the structure to which they are attached will be addressed. ‘The various test methods for the assessment of the proper functioning (suitability) of fastening systems are reviewed by the appropriate regulatory bodies, e.g. UEAtc, EOTA and ASTM. 2. Terminology 2.1, Introduction ” In the following sections the terminology is presented in two forms (@ CEB-preferred definitions and notation; this notation system, although originally developed for the CEB/FIP Model Code, has become the basis for ISO 389; itis adopted here as a uniform system of notation (b) common-usage notation from various countries and organizations; this has been included to enable the reader to identify the terminology contained in the references, and to allow direct comparison with the CEB-preferred terminology. The following organizations are referred to by acronym. ACI Afnor Au ASTM CEB CEO EOTA TEBt Iso PCL UEAte American Concrete Institute Association Frangaise de Normalisation Architectural Institute of Japan American Society for Testing and Materials Comité Euro-International du Béton (Euro-International Concrete Committee) Comité Européen de ’outillage (European Tool Committee) European Organization for Technical Approval Institut flr Bantechnik (Institute for Construction Technology) International Organization for Standardization Prestressed Concrete Institute Union Européenne pour I’Agrément Technique dans la Construction (Buropean Union of Agrément) Fig. 2.1 (let). Types of loading: (a) tension loading; (b) compression loading; (c) shear loading, (d) combined tension and shear loading; (¢) combined tension and shear loading with bending Fig. 2.2. Components of a connection 2. Fastener |. eure tl 4. Thabtoning torque 4 2 Prestress force 3. Base material 43. Olamping force 4 Expansion force 5. Friction force TERMINOLOGY Fig. 2.4. Mechanisms and L ‘components of displacement A ‘of an expansion anchor Displdcement 722] Elongation = movement ofthe anchor contibuted by Steel strain and slastc conerste strain Sip= or-ecovarabe @.| movement of te | Unloaded ond of a fastener Displacement = Alp + elongation +. Expanded pastion 2. Balt pulkout (ro sp of sleeve) undor oad 3, Sip of sleeve under Toad—ne batt puFout ope 8,2,.0.9 $e ney Se! of installation: (a) flush inscallation, pre-positioned Jastening; (b) flush h Installation, in-place i fastening; (c) stand-off : installation L Fig. 2.5. Types of Fig. 2.6. Expansion and spliting force: (a) expansion pressure forp 0) splitting force Fy = Lfegsind dé @ » : | cos INTRODUCTION Table 2.1 (below and facing page). Summary of terminology Preferred terms Alternative terms Definition Admissible load - Load allowed for a fastening under service conditions per a design code Anchorage depth Embedment depth _| Effective depth of the fastener (sce Fig. 2.7) ‘Anchor loading — axial | Pull-out Load application is in the direction of the axis of the (tension or compression, | Tensile load fastener see Fig, 2.1(a) and (b)) | Tension Compression ‘Anchor loading — bending Shear or combined loading applied to an aachor in Gee Fig. 2.10) such a manner as to induce flexure in the anchor body ‘Anchor loading — Oblique tension Combination of tensile and shear loading applied combined (see Fig. 2.1(8)) simultaneously Anchor loading — shear} Lateral Direction of load application is perpendicular to the (see Fig. 2.10) axis of the fastener Base material (see Fig. 2.2) | Anchorage material | Material into which a fastener is anchored, e.g. ‘Structural component | concrete, masonry Structural unit Support Clamping force (see Fig. ‘Compressive force between the base material and the 23) fixure Concrete (see CEB-FIP ‘Antficial stone formed by mixing cement, aggregate, Model Code, 1990) ‘admixtures and water: detailed definitions are given in the CEB—FIP Model Code (1990) for concrete structures Connection (see Fig. 2.2) | Anchorage assembly | Combined assemblage of fixture, fastener and base ‘Anchor fixing material Fastening assembly Diamond drill bit - A dill ti with a diamond matrix Displacement (see Fig. 2.4) | Movement Movement of the exposed end of a fastener under load relative to the free surface of the base material outside the zone of influence of a failure cone Drill Drilling toot Electric or air-powered tool for making holes: drilling is characterized by rotary action, often supplemented by percussive or hammering blows to promote rapid hole formation Drill bit ‘Tool for forming a hole Elongation (see Fig. 2.4) | — ‘Movement of the anchor under load contributed by } strain of steel and elastic concrete strain (excluding slip) Expansion force (see Fig. | Spreading force Force induced in the base material by expansion 2.6) mechanism or anchor head, normally radially symmetrical to anchor Failure mode Failure mechanism of a fastening under load Fastener (see Fig. 2.2) Anchor Element used to fasten a fixture to the base material (see Fig. 2.2) Fastening system Anchor system Fastener and corresponding drilling and setting tools TERMINOLOGY Preferred terms Alternative terms Definition Fixture (see Fig. 2.2) Assembly unit ‘The building component attached to the base material ‘Component to be fixed | by a fastener (see Fig. 2.2) Attachment Flush installation (see Fig. | — Installation in which the fixture is in contact with the i 2.5) surface of the base material 3 Follow-up expansion Re-expansion (German | In an expansion anchor, displacement of the cone oF } Nachspreizen) mandrel under load which serves to further expand the spreading elements of the anchor, accompanied by little or no stip : Friction (see Fig. 2.3) - ‘A load-transfer mechanism proportional to the force applied normal to the friction surfaces Grout - Fluid mixture of cement and water, possibly including fine aggregate and/or admixtures Invplace fastening (see Fig. | Through-assembly Applicable only to certain types of expansion, 2.5) (German undercut, grouted and chemical anchors, which can be Durchsteckmontage) _| inserted through the fixture into the hole : Masonry - An assembly of prefabricated units stacked and bonded together with mortar. Hollow assemblies may or may not be filled with grout. Masonry units may be made from clay (clay, loam with brickearth or shale); dense, lightweight or autoclaved aerated concrete; calcium silicate or natural stone. Units may be solid or may contain holes or cavities. Mortar (German Mariel) ‘Any cementitious or polymer-based compound used to bond an anchor to a base material g Pre-positioned fastening = | — ‘A connection where the fastener is installed before the : (ee Fig. 2.5) fixture is attached Prestress force (see Fig. | — ‘The axial force induced in a fastener during torquing 2.3) of the bolt Recommended load Maximum, ‘Maximum service load recommended by a | recommended load | manufacturer: this load value may differ from the admissible load Relaxation (see section 5.1) | — Reduction in clamping and prestress force with elapsed time Slip (see Fig. 2.4) - Non-recoverable movernent of the unloaded end of a fastener under load relative to the base material: does not include steels strain or elastic concrete strain, : Spliting force (see Fig. 7 ‘Component of spreading force which acts to split the & 26) base material : Stand-off installation (see | Spaced installation | An anchorage assembly in which the fixture is secured : Fig. 2.5) German at a distance from the surface of the base meterial i Abstandsbefestigung) & ‘Stop drill a Drill bit equipped with a drill stop that ensures a z precise predetermined hole depth g Tightening torque ad ‘Torque applied to a fastener, preferably by means of a : torquemeter during installation Ultimate load Capacity Maximum load sustained by a connection prior to [ fiture Table 2.2. Notation: dimensions Dan ex ¢50, [ceo] franc] cera] su [Oe [ne] seoten_[uear]usa SS |°°° ee ees | Ser [eta [ Soren] [exc eee] fee ele OP oreo] (ea ‘report 815- 5080) ) Bre | its 2.7) Depth Pa acta eat ttt) Ae et | tet a ttt Eee lneberrnereny I os nen te elude weecetscenw fer do ty ik Bi sete ar ‘ ing Sr age tomer | Le I Ceet eects fia Hla Tle a (Gixtuce) fee id, | dy 4, hk ty d, rt Tee ot moan ter |e ‘ é Tae face ae Mn mae » if le o | i 5 ee ef sande pao — ta | | Diner a aso cin dence |d a lam tee we sce f ete eel oly (ft |e A eof tart (Sy a ie |e pe ie iB er a re Thence ort sa | | fectacees [a ae ja de ln | eee traits cue (ez jas (Sew (d a Ie rae aretha (ar weet (e é Scand eigen Se ae ee nen ea a ae fates cele let St ac es satire! 7 ae See Cae enor | See ta eat a Bs Coa ce of taxes in group | Bee la Oe a eee eI ee tt etna tata ae se, Cee tees es | : Si ee nls’ fez (a a loz Sant see oes on ey 7 her dare : : 7 aa T Be tere vn 4 7 ae os Mecote: nese ara | A ala Fan td hy, may be the same Tor some fasteners, +f For plastic fastener only Usually dng do IF Spacing required for maximum anchor capacity, § Euge distance required for maximum anchor capacity. Table 2.3. Notation: quantities* Desccption [CEB (80, [CEO] France| Germany| res: [Grest [Sepa] Sweden | UBAte| USA 398) used | |(Afnor| (8, | Britain) Britain] (AU |(Boverker) {ACD in this: (E27. | (BS |(CFA) 19771) Canada report (sec BIS: 5080) Fig. 2.7) | 816) Loads, forces | ‘aia oad ot force (ension or E ‘compression)t N A IN N : Bending moment M a Mu ‘Clamping force Fa Expansion force Fey | Internal bolt force Fe Pretess force P Shear load or forcet v Fe v v Spliting force Fe Tensile fur fad in test (Gngle value) % FR. |Xin [Fe Me Mean tensile fare Toad in a test, series Nm = ye Pow Tightening torque Tee My Mu Torsiont t r Admissible tension load Nsw X_ | zul F N, Celealated design tensile capacity | Ne ’, Pe Recommented tensile led Nee [rook oN, Sensis Concrete compressive strength bated on cylinders lé Concrete compressive strength, hractoets value (ere. 3% Face) lee lia concrete compressive strength, design bes Mes le : concrete tesile strength le ey Hi Modulus of elasiity é See tense strength A Mie soe |e lhe Stee yield strength ie le sop eos Sess Aree 4 ‘Aes of concrete ae Siresed ares of tel a 4, 4 a Sree srese o fe Admissible ste stress sun ou F, Bond tess ne es n Coefficients Satay factor (or coeftcien) fy Recution factor $ x x i Redvetion fctor for eracked % conerete Yn Xn a Reduction factor due to spacing | ¥, x & Redstion factor de t0 distance to * free edge % Xe s Reduction factor due to eooentisty | ¥. xe : Reduction factor dus to thickness of| e ‘structural component ty xa : Swength redoetion factor ° | $ Statistical analyse Number of fasones in a test series * * * a [ns |e Value of individual est eesuit | % Xi fs x ik |k Aithmetic mean value \: x fe x Standard cevation 5 5 : 5 Pee HeeeL Coeticet of variation 5 * * ; ' soi Subscripts: adm admissible; ¢ concrete; (cylinder strength; jee concrete compression strength measured on cubes fu.20 ‘concrete compression strength measured on cubes of side 200 mm), d design; m average; rec recommended; § steel; L ulna E + Load in general sense of an action effect 5; load in general sense of resistance R; S and R may be used as subscripts 10 : denote te nature of load & £ In this report “concrete tensile strength’ refers to splitting tensile strength, noise ownuog | STULL A Soni it I} @ ean rove welepun ce oe oe pamunnnben : seus 0 mow SE sioqpqvoo-wonsintyeg s ; wore FS: rou pepo your pope fF A {eve msiepun, INTRODUCTION + at TERMINOLOGY Fig. 2.7 facing page). Dimensional nomenclature for principal fastener types (details are given in chapter 3) {) general; (b) positioning in base material; (c) thickness; (d) drill bit Fig. 2.8 (below). Non-prestressible anchor-assemblies (no clamping force) @ 2.2. Glossary of terms 2.3. Notation h © @ Some of the terms are international in usage; others have alternatives, which are listed where appropriate. Where a suitable English term is not available to describe a particular object, action or phenomenon that is referred to frequently in the text, an ‘invented’ shorthand term has been adopted. This is accompanied by the original non-English term in parentheses under “Alternative terms’ (Table 2.1). Tn Figs 2.22.8 the anchor features may be exaggerated to illustrate force transfer mechanisms; no similarity to a particular anchor type is intended unless specifically stated, Dimensional notation used in this report is described in Table 2.2, with reference to Fig. 2.7. In Fig. 2.7 expansion anchors are shown in the unexpanded position. The notation agrees with the CEB—FIP Model Codes (1978, 1990). Quantitative notation used in this report is described in Table 2.3. 3. Types of fastener Fastening and connection systems must typically satisfy various performance criteria, including one or more of the following. @ Strength; the capacity of the fastening to resist the forces to which it will be subjected during its lifetime, including those caused by external loads and restraint of imposed deformations (e.g. due to temperature variations) and those required to maintain stability © Ductility: strictly, the ability of a fastening to accommodate relatively large inelastic deformations without a significant decrease in capacity. In seismic applications this characteristic may be enhanced sufficiently for large anchorages to provide some degree of energy absorption capability. Its also a measure of a fastening system’s capacity to sustain overloads without precipitous strength loss. (© Durability: the resistance of a fastening system to the adverse effects of variations in temperature and exposure to moisture or other corrosive agents. Such design factors as cost, constructibility, maintainability and appearance ‘may also figure significantly in the final design of fastening systems. The inability of any one system to satisfy all these criteria equally well for all applications has given tise to a multitude of fastening types. The load-transfer mechanism employed by the fastening system usually determines many of the performance characteristics discussed above. An understanding of the principal load-transfer mechanisms across the interface of the base material and the fastening is thus necessary for the classification of fastening systems. Three different types of load-transfer may be identified for tension loading (Fig. 3.1(a)—(©)) mechanical interlock, friction and bond. ‘The term ‘bond’ as used here refers only to the continuous transfer of load along the embedded length of an anchorage. A further distinction may be made between mechanical and chemical bond, as in the case of chemical and grouted anchors: since the transition between these two load-transfer mechanisms is generally difficult to establish even at a microscopic level, the traditional and broader definition of the term ‘bond’ as explained above has been adopted unless otherwise noted in the text. ‘Mechanical interlock is thus distinguished from bond in that the load transfer is concentrated in one region, usually near the end of the anchorage, and results in locally high bearing stresses. In the case of friction, load transfer is again not defined at the microscopic level; for the purposes of this report the identifying characteristic of this load-transfer mechanism is the proportionality of the load transfer to the force applied normal to the friction surfaces. Externally applied tension loads may be transferred to the base material by a single mechanism or by a combination of mechanisms. In all cases, the surrounding base material is subjected to both compressive and tensile stresses resulting from the force introduced by the anchorage. In the case of shear loading, the applied force is resisted by a compression 4} v force couple generated by bearing of the anchor body on the base material need (ee Fig. 3.1(@)). The distribution of compressive stresses depends on the a relative stiffness of the anchor body in flexure and the base material in compression, and generally changes with increasing load. Simultaneous Fig. Pe 1. Load-transfer flexural and catenary tension forces are developed in the anchor body with feat nee imeriock; inereasing shear loading. As the base material at the surface of the anchorage (0) tension load, friction; i$ progressively crushed, the resultant of the upper compression stress block (0) tension load, bond, moves away from the applied shear force, thus increasing the applied moment (6) shear load ‘on the anchorage. With pronounced bending of the anchor body, the catenary 0 3.1. Cast-in-place systems Fig. 3.2. Threaded inserts: {a) loop: () deformed Section; (¢) anchorage member; (d) bolt head TYPES OF FASTENER tension force becomes significant and may reach up to 40% of the shear force (Fuchs, 1990). This force is resisted by a combination of the mechanisms discussed above for tension loading. ‘The anchoring systems currently in use may be classified into two broad categories: cast-in-place systems which are positioned in the formwork before the concrete is cast, and post-installed systems which are installed in hardened masonry and concrete structures, ‘The anchors discussed in sections 3.1 and 3.2 are typically intended for use in concrete; masonry anchors are described in section 3.3. ‘The most common types of cast-in-place anchors are headed bolts, J-bolts or L-bolts, threaded rods and reinforcing bars. All these embedments may or may not have an end anchorage such as a loop, head, nut and/or plate washer to enhance mechanical interlock. Other forms of prefabricated cast- in-place anchors include threaded inserts, welded studs, deformed bars, channel bars and shear lugs. Cast-in-place embedments are available in a variety of configurations and lengths. So-called ductile embedments are designed to develop concrete pull-out and shear strengths that exceed the capacity of the steel bolt in the connection. The degree of ductility associated with such an embedment depends on the mechanical properties of the bolt and the length over which inelastic deformations can occur. ‘Cast-in-place systems transfer the load into the base material by mechanical interlock (bearing stresses) and/or bond. In concrete construction, they are typically positioned in the formwork and may thus be used even in highly reinforced members without the difficulties associated with post-installed systems. Also, the reinforcement in the vicinity of the anchorage may be sized and configured to carry more effectively the loads introduced by the anchorage into the structure. The extent to which these types of load can bbe accommodated in the design of the reinforcement is of course determined by the availability of information on the location and magnitude of the loads associated with the anchorage during the design process. 3.1.1. Mechanical interlock systems LLL. Threaded inserts. Threaded inserts such as those shown in Fig. 3.2 are typically manufactured from wire, rod or cast steel which is internally threaded, and may consist of a loop (Fig. 3.2(a)), a deformed section (Fig. 3.2(b)), a spiral, J, L or straight anchorage member (Fig, 3.2(c)) or @ bolt head (Fig. 3.2(@). These anchors are available in a wide range of sizes, and ate often used in primary structural connections in precast structures as well as for lifting precast members such as walls, beams, columns and piles. «) ® ic) @ INTRODUCTION Fig. 3.3. Bolts of various ‘shapes: (a) headed bot, (©) L-bolt; (c) J-bolt; (d) headed stud 2 ‘They may be threaded to accept threaded rods, standard bolts or machine bolts; they are also manufactured with coils to receive lifting Ings. Inserts are usually set flush with the concrete surface, and are positioned in the formwork by use of templates or wire ties. ‘The inserts shown in Fig. 3.2(b)—(€) provide mechanical interlock with the concrete; the insert shown in Fig. 3.2(a) relies on bond between the reinforcing bar and the concrete. 3.1.1.2. Headed bolts. Headed bolts (see Fig. 3.3(a)) are standard structural steel bolts placed with the head embedded in the concrete. Such an anchorage ‘depends primarily on mechanical interlock (bearing) at the bolt head for tension load transfer, although some bond along the bolt length may also be developed. ‘Typically, bolts used for anchorage purposes are made of structural grade Jow-carbon (mild) steel with coarse rolled threads. An over-sized washer fabricated from plate steel is often used to increass the bearing area at the bolt head. Headed bolts are often installed in fresh concrete by pushing or tapping of:the bolt into place and vibration of the conerete to consolidate it around the bolt shank, This practice often results in voids in the concrete surrounding the bolt head, and is generally discouraged. More typically, headed bolts are placed in the formwork by means of templates or wire ties 3.1.1.3. L-bolts and J-bolts. L-bolts (see Fig. 3.3(6)) and J-bolts (see Fig, 3.3(@)) are typically manufactured from smooth bars. The bolts are bent into aJor L shape at one end and are threaded at the otter end with coarse rolled threads. Together with headed bolts, hooked bolts are typically used in foundations. Tension loads are transferred into the base material primarily by mechanical interlock. 3.1.1.4, Threaded rods and hooked reinforcing bars. Standard threaded rods may be bent at one end (cf. Fig. 3.3.(b) and (¢)) or used in conjunction with a washer and hexagonal nut (cf. Fig. 3.3(e)). They are typically used as hanger rods, shear studs or dowels in concrete construction. Threaded rods are readily available in various diameters and lengths, and may, in the case of partially threaded rods, be manufactured with up-set threads to provide an equivalent tensile area over the threaded length. They are generally available in the same variety of steel grades and types as structural grade bolts. Hooked reinforcing bars may be threaded at one end to provide an anchorage similar to-a thre&ded rod. Right-angle or 180° hooks are typical. Tension loads are transferred into the concrete by a combination of bond along the bar length-and mechanical interlock at the hook. 3.1.1.5. Welded concrete anchors (headed studs). Headed studs (see Fig. 3.3(@)) are used for all types of connection in various applications. The stud ILGT @ ) Fig. 3.4. Channel bars Fig. 3.5. Bonded anchorage (deformed bars or threaded rods) TYPES OF FASTENER 2 |p s0-L meter [Channel bar (Channel bolt Nut rrure comprises a smooth shaft with a cold-formed round head. A ratio of head diameter to shaft diameter of ~ 1.5 is typical. For anchorage purposes, studs may be welded to embedment plates or angles by use of the electric-arc stud welding process, or threaded at the free end and used in the same manner 1s the headed bolt (see section 3.1.1.2). Headed studs transfer tension loads to the concrete by mechanical interlock. 3.1.1.6. Shear lugs, Shear lugs are steel plates, usually welded to the botom of column bases, which transfer shear load into the concrete through bond and bearing. The behaviour of shear lugs belongs to the more general topic of building foundations, and is not covered further in this report. 3.1.1.7. Channel bars. Channel bars (see Fig. 3.4) comprise cold-drawn or hot-rolled U-shaped steel profiles with special anchoring elements attached to the embedded side. They may be either prefabricated or designed for specific applications. The channels are typically filled with rigid foam (e.g. Styrofoam) to facilitate attachment of the channel to the formwork and to prevent intrusion ‘of concrete into the channel during the concrete pour. After removal of the formwork the rigid foam is extracted from the channel with a knife, and the bars are attached to the channel with special hammer-head or hook-head bolts or lugs. ‘The anchoring elements attached to the back of the channel may comprise ‘T-shaped anchors or headed bolts welded or forged to the anchor channel. 3.1.2. Bonded anchorages In the case of straight threaded rods and deformed bars embedded without supplemental end anchorage as provided by a hook or washer and nut (see ‘Stool pate ond stress. 2 Reinforcing bar (wetdable steel) INTRODUCTION 3.2. Post-installed systems Fig. 3.6, Classification of expansion anchors Expansion anchors “ ‘section 3.1,1.4), tension load transfer is accomplished by bond alone. Straight bars and rods are commonly used in the same manner as hooked bars and rods, and are distinguished only by the absence of the mechanical interlock load-transfer mechanism (Fig. 3.5). Post installed systems may either be installed into pre-formed or drilled holes, or be driven into the hardened base material with impact energy (as in the case of direct fastening systems). 3.2.1. Drilling techniques The most common method used to drill holes in concrete and masonry is combination of rotary drilling and percussive action as generated by the rotary percussive drill. By use of carbide steel helical bits, holes up to 30 mm in diameter can easily be produced with hammer drills. Small-diameter reinforcing steel in the concrete may be broken with a large drill, and even incidental contact between the drill bit and reinforcing bars may result in damage to the reinforcement. ‘A recent improvement in the design of rotary percussive drills is a built-in vacuum to facilitate removal ofthe drilling debris (ground concrete) generated during the drilling process. This allows a considerable reduction of dust levels, shorter drilling times and increased rill bit lite. Diamond drills are required for many applications involving larger and/or deeper anchorages. Holes of almost any length in diameters up to ~200 mm can be drilled with & combination of bits and extenders usually powered by a high-torque low-velocity electric motor. The end of the cylindrical hollow- core drill bit is usually coated with a diamond matrix set in an adhesive'base. Single-cone type (Fi. .7(a)) Double-cone ype (Fig. 3.7(6)) Bieave ype aod Taper-bolt type (ig. 8.7) ‘Torque-contoll expansion anchor Wedge type (Fig 8.710) Botips rt Bot with intornaly: ‘threaded cone (Fi. 9.72)) Cone-down type (Fig. 2.9(a)) opin anchor Wedge-down ype aod Shank-down type (Fig. 3.9(0)) Deformation-centolled expansion anchor Gone type Fig. 3.916)) Selecting anchor (Fig. 3.910) Lead caulking anchor (Fig. 3 9(6)) ‘Shal-down pe thd Plug not type (Fig. 3900) Fig. 3.7. Expansion ‘mechanisms of various torque-controlled expansion ‘anchors: (a) single-cone type; (b) double-cone type; (6) taper-bolt type; (@) wedge type; (@) bolt with internally threaded cone ‘TYPES OF FASTENER In order to preveiit overheating of the adhesive and to assist in the removal of drilling flour from the hole, water is typically pumped through the core bit and circulates up the sides of the bit and out of the hole during drilling. Diamond drilling without water is possible but slower, and results in increased bit wear and drilling vibration. Diamond drilling gives a smoother and ‘geometrically more precise hole than rotary percussive drilling, and diamond bits can cut through large reinforcing bars without significant deviation from the drilling path. Considerable skill is generally required of the drill operator to prevent the accidental destruction of reinforcing. steel during drilling. - 3.2.2. Expansion anchors Expansion anchors differ from other types of post-installed anchors in that they transfer tension loads to the concrete principally by friction. The frictional resistance of the anchor depends directly on the normal forces generated by the anchor expansion mechanifm during installation and throughout the life of the anchorage. Expansion anchors can be categorized into two general classes, according to the manner in which the expansion forces are generated (Gee Fig. 3.6). Torque-controlled and deformation-controlied anchors may be further subdivided according to geometrical characteristics and methods of installation respectively. ‘Torque-controlled expansion anchors are expanded by the application of a specified torque to the bolt head or nut. The applied torque serves to draw a conical wedge between the spreading elements, forcing them outwards against the sides of the hole in the concrete. The concrete deformation (expansion deformation) depends on the prestressing force and the deform- ability of the concrete. Tension loads are delivered from the bolt through INTRODUCTION the wedge and into the spreading elements, where they are transferred by friction to the concrete,; External loads serve to force the wedge tighter between the spreading elements, thus increasing the spreading force. Deformation-controlled anchors differ from torque-controlled anchors in that the conical wedge is driven between the spreading elements of the anchor with impact energy, usually produced with a hammer and a setting tool. The amount of spreading force generatéd depends directly on the extent to which the wedge is initially driven between the expansion elements. In most cases, the external loads on the anchor are, delivered from the anchor bolt to the concrete directly by way of the anchor shell. Since the external load does not pass through the wedge, the amount of spreading force generated by these anchors is maximum at the time of installation and is unaffected by the external loads applied to the anchors during their life. Thus, deformation-controlled ‘expansion anchors, which cannot undergo further expansion, are sensitive to the dimensional inaccuracies of the drilled hole. 3.2.2.1. Torque-controlled expansion anchors. The many types of torque- controlled expansion anchors available can be classified as either sleeve-type ‘or bolt-type, depending on the relative diameter of the bolt and the drilled hole (see Figs 3.6 and 3.7). Bolt-type anchors are recessed at the expansion ‘end to accommodate the spreading elements within the diameter of the bolt. Thus the diameter of the drilled hole is nominally the same as that of the bolt. In the case of sleeve anchors, the diameter of the drilled hole must be equal to the diameter of the sleeve that encases the bolt. Often the design of the anchor calls for the sleeve to project beyond the surface of the concrete to engage the part being fastened for shear loads. Sleeve-type anchors are available in many configurations, which vary with respect to the number of spreading elements or cones and the length and construction of the sleeve. Taper-bolt sleeve anchors (Fig. 3.7(c)) differ from cone-nut type anchors in that the conical wedge is machined into the end of the bolt much like a bolt-type anchor, whereas cone-nut type sleeve anchors (Fig. 3.7(a) and (b)) employ internally-threaded conical elements. The most common form of bolt-type anchor is the wedge anchor (Fig. 3.7(d)). A typical wedge anchor assembly-comprises a steel bolt, expansion elements and a nut and washer. The bolt is generally coarsely threaded at ‘one end, and has one'to two uniformly tapered mandrels machined into the embedded end. The’expansion element typically consists of a spring steel collar or multiple collar segments, which before installation of the anchor may be secured loosely to the mandrel with wire or tape. The expansion clement typically has deformations for engaging the side of the hole during installation A hybrid of the bolt- and sleeve-type torque-controlled expansion anchors is shown in Fig. 3.7(e). The upper part of the anchor consists of a bolt which is internally threaded at the embedded end to accept a lower assembly consisting of a threaded bolt with a mandrel and an expansion sleove. Figure 3.8 shows the elastic stress distribution generated in PMMA (Plexiglas) by a typical sleeve-type expansion anchor. Application of torque generates a prestressing force in the bolt which is counterbalanced by compression in the base material. The expansion forces generated by the anchor likewise induce high compressive and tensile stresses in the base material. In concrete, the stresses induced by expansion of the anchor greatly exceed the elastic capacity of the concrete. Local crushing or plastification allows the.stresses to be redistributed over a larger area of the concrete surrounding the expansion end of the anchor. ‘The prestressing force and the spreading force decrease due to relaxation of the concrete and other effects over time. If the externally applied tension load exceeds the remaining prestressing force, the cone is drawn tighter between the expansion elements and generates the higher spreading forces Fig. 3.8. Distribution of principal elastic siresses in PMMA (Plexiglas) generated by a torque ‘conrolled sleeve-type ‘expansion anchor (ken from Seghezzi (1984)) PYPES OF FASTENER necessary to resist the applied load. Further expansion is possible only if the frictional resistance between the cone/mandrel and expansion elements is less than the friction generated between the expansion elements and the surface of the drilled hole: otherwise the anchor may be pulled out with no further increase in load beyond the prestressing force level. ‘As noted above, torque-controled expansion anchors are set with the application of a prescribed torque. The ability of the anchor to develop this, torque without turning in the hole serves as a control for correct installation. ‘Anchors are typically considered to be properly installed only ifthe required torque can be achieved during installation. Ta addition to the frictional force-transfer mechanism described above, torque-controlled expansion anchors may transfer load to the concrets by ‘means of mechanical interlock causéd by inelastic concrete deformations at the expansion end of the anchor. This mechanism, however, is generally thought to represent only a small part of the load capacity of most torque- controlled expansion anchors. 3.2.2.2. Deformation-controlled expansion anchors. These can be classified as either wedge-down or shell-down (Figs 3.6 and 3.9). Wedge-down deformation-controlled expansion anchors comprise an internally threaded shell and an expansion plug (Fig. 3.9(a)) or a hollow bolt and drive pin (Fig. 3.9(0)). These anchors are set by driving of the plug or pin into the expandable end of the anchor with a hammer or setting tool. ‘The anchor shown in Fig. 3.9(a) is commonly known as a drop-in anchor. Shell-down deformation-controlled expansion anchors comprise an expandable sleeve and an expansion plug (Fig. 3.9(c) and (d)) or bolt with ‘a machined mandrel at one end (Fig. 3.9(€)). They are anchored by driving of the sleeve with the help of a hammer drill onto the cone, causing the expansion sleeve to spread and press against the wall of the pre-drilled hole. ‘The self-drill anchor is a special form of shell-down anchor that is intended to climinate the problem of mismatch of anchor and hole diameter associated v INTRODUCTION Fig. 3.9. Design and ‘operation modes of metal deformation-controlled expansion anchors: (@) cone-doun type (‘drop- in anchor’); (b) shank-down ‘ype (‘stud anchor’); (©) cone type; (d) cone type (Cselfdril anchor’); (@) lead caulking anchor; (9 plug- bolt ype < t © © a with the drop-in anchor. The shell is chucked into the drill and used to drill the hole in the concrete. The conical plug is then inserted into the end of the anchor, and the anchor shell is driven over the plug using the hammering action of the rotary percussive drill. This design places conflicting demands ‘on the material used for the drilling and expansion end of the shell. While the steel must be hardened to serve as a drill bit, it must also retain sufficient elasticity and toughness to sustain the deformations associated with the ‘expansion process. A lead caulking anchor is shown in Fig. 3.9(f). The anchor is set by use of a setting tool and a hammer to force the lead sleeve over the expansion cone. There are two types of Iead caulking anchors: those that are installed individually as a single unit, and those that are installed with two or more sleeves (generally referred to as multiple or compounded units). The single unit anchor generally has a longer threaded steel cone. The compounded anchor is assembled from multiple lead sleeves in combination with either threaded or unthreaded expansion cones. A bolt may be inserted through the assembly so that the bolt head bears on the Test expansion cone, in which case the expansion cones need not be threaded. The installation instructions typically require that each lead sleeve be expanded separately with the setting tool and hammer Deformation-controlled expansion anchors are set by the expansion of the sleeve to a certain expansion ‘deformation’ which leads to the formation of an expansion imprint in the sides ofthe hole wall. The anchor bearing capacity depends on the degree of concrete deformation achieved during expansion, indicating that the load-transfer mechanism is a combination of mechanical interlock at the expansion imprint and ftiction. The free play between anchor ‘and wall of the drilled hole and the deformability of the concrete determine the extent to which an expansion imprint may be formed. ‘ y ‘TYPES OF FASTENER Ifinstalled correctty, deformation-controlled anchors generate a considerably higher expansion force than torque-controlled expansion anchors of similar size (Mayer, 1990). As discussed above, this force is diminished over time by relaxation of concrete. As explained in section 3.2.2, the anchors shown in Fig, 3.9(a)—(©) are incapable of generating increased spreading force in response to externally applied loads. As the anchors cannot expand further, their load-bearing behaviour depends on the extent to which expansion of the anchor was achieved during installation. Since the maximam ‘deformation’ of the concrete at the expansion end of the anchor is constant for a given anchor type, anchor performance depends on the dimensional accuracy of «the drilled hole. If the drilled hole is too narrow, the expansion force during ‘setting of the anchor becomes very high, with the risk that anchors set too close to an edge or to each other will spall or split the concrete. In high- strength concrete, this anchor type can be difficult to set properly and an imprint of the spreading elements is formed in the concrete to a lesser extent. Due to the sensitivity of deformation-controlled expansion anchors to dimensional inaccuracies of the drilled hole, the use of drill bits of correct nominal size and within the prescribed tolerances is very important, and inspection of installation is indispensable. In the case of drop-in anchors (Fig. 3.9(a)), a visual check for full expansion may be made by use of the setting tool. This assures the correct position of the expansion plug flush with the end of the shell in the drilled hole. For the anchors shown in Fig. 3.9(c) and (d), the distance from the top surface of the expansion plug or cone to the end of the sleeve should be measured and compared with the installation specifications for the anchor. ‘The anchor shown in Fig, 3.9(f) transmits the load to the concrete by way of the expansion mandrel at the end of the bolt, Unlike other types of deformation-controlled anchor, this configuration allows for continued expansion with increasing applied load. 3.2.3, Undercut anchors By definition, undereut anchors transfer tension Toads to the concrete primarily through mechanical interlock. Although other types of expansion anchor, notably self-drill anchors, may develop some mechanical interlock with the concrete, the degree to which the undercut anchor depends on this load-transfer mechanism distinguishes it from other anchor types. ‘All undereut anchor systems require a hole in the base material (concrete) that is locally widened or ‘undercut to accept the expansion elements of the anchor. Depending on the system, this special hole geometry may be created before or during setting of the anchor. Since they rely on mechanical interlock for load transfer, undercut anchors typically develop little or no expansion force in the concrete during installation. As a result, the tensile stresses present in the concrete during prestressing and loading of the anchor are considerably ower than for metal expansion anchors. ‘The undercut geometries created by various undercut anchor systems are shown in Fig. 3.10. These can be generally classified according to whether the undercut widens at the top (Fig. 3.10(a)) or at the bottom (Fig. 3.10(6)— (©). A distinction may also be made based on the stage at which the undercut is formed during the installation process. For the anchors shown in Fig. 3.10(a)—(6), the undercuts are formed by special drilling tools before anchor installation. The drilling tools typically use either a carbide-tipped bit or a water-cooled diamond drill bit. Often the undereut is formed during a second Grilling step after the hole is drilled. The undercut of the anchors as shown in Fig. 3.10(¢) and (2) is formed during anchor installation by driving the diamond. or carbide-tipped expansion sleeve over the expansion cone either by rotation and hammering (Fig. 3.10(6)) or by hammering alone (Fig. 3.10(¢)). Again, the degree to which these anchors develop a mechanical 19 INTRODUCTION interlock with the base material distinguishes them from the anchors shown in Fig. 3.9(c)—(f). Capsule-type bonded anchors set in undercut holes are available (see section 3.2.4.1). 5.2.4. Bonded anchors ‘The term “bonded anchor’ as used in this report encompasses the entire range of mortar-filled embedments, including cementitious and polymeric-based systems (see Fig. 3.11). Bonded anchors are further subdivided, according to the method of placement, as capsule-type ot injection-type. ‘The typical.capsule-type anchor comprises a cylindrical glass ampule containing an uncured polymeric compound and, in isolated form, an accelerator or catalyst. It is designed to be inserted in the hole before setting of the anchor rod. Mixing and (in the case of resin-based systems) curing ‘occur during the process of drilling through the capsule with the anchor rod. The origins of the capsule anchor can be traced to rock-anchor systems employing resin-filled polyethylene bags. Injection-type anchors include the entire range of systems where the mortar compound is introduced into the hole in an unpackaged form. The mortar may be polymeric-based or cementitious in composition. Typically, but not invariably, the anchor rod is placed in the hole before introduction of the grout. The method of installation generally influences the load-bearing behaviour significantly (particularly with respect to cleaning of the drilled hole). In this report, anchors that are bonded by a poured-in (as opposed to injected) grout are called ‘grouted anchors’, and are classified as a subset of injection anchors (see section 3.2.4.3) ‘The resins employed for bonded anchors are based on a range of polymeric compounds, including unsaturated polyester resins, epoxy acrylate resins (vinyl ester), and pure epoxide resins (epoxies). The polymer components are packaged in four main forms (a) glass capsules (®) tubes (0) plastic cartridges (@) bulk packaging of components (e.g. barrels, cartons, cans). ‘While (4) requires proportioning and mixing of the polymer components on site, (a) —(6) deliver pre-measured amounts of resin and hardener or catalyst ‘0 as to assure automatic and (celatively) consistent proportioning and mixing. Polymer adhesives can be formulated to have either fast or slow setting ‘times (short or long pot life), depending on the requirernents of the application. Except for polyesters and vinyl esters, a significant length of time (curing time) must elapse after the adhesive has set before the anchor can sustain Toad, and in all cases premature loading or disturbance of the polymer matrix will preclade the attainment of full strength. All polymer compounds have a limited shelf life, and must be stored away from sunlight-and extreme ‘temperatures. The load-transfer mechanism primacily associated with bonded anchors is the bond between the anchor rod and mortar and between the mortar and the wall of the drilled hole. In general, deformed bars or threaded rods are used to ensure adequate bonding between the mortar and the anchored element. Inregulatities in the sides of the hole may improve the bond between mortar ‘and base material, and thus add to the total strength of the anchorage. The presence of drilling dust and fragmented base material on the sides of the hole may significantly impair the formation of bond between mortar and base ‘material Recently, bonded anchors set in an undercut hole have been developed, ‘These transfer a tension load by bond through the anchor rod into the mortar, and mainly by mechanical interlock from the mortar into the concrete. Fig. 3.10 (above). Typical iidercur anchors: (a)—~(c) undercut is formed before anchor installation; (d) and (6) undercut is formed during anchor installation carbide driing gin ‘Bonded anchors Fig. 3.11 (right). Classification of bonded ‘anchors TYPES OP FASTENER © Poiyester Eporyacryiate ‘Wing! aster) Epoxy Polyester Epoxyacryate (uiny esto Epoxy [Comenttious (routed anchor) ‘others a INTRODUCTION Embodment ‘Threaded rede with nut depth mark Capsule with resin, Resin, ardering agent Sendnng agent and ‘and quart aggregate quartz aggregate o o Fig. 3.12. Copsule-type Unless an expansive grout is used, no expansion forces are generated during anchor: (a) instlaion: @) setting of the anchor; however, expansion forces are created during installed anchor prestressing and loading of the anchor due to formation of hoop stresses in, the base material. These expansion forces are considerably lower than those generated by friction-based expansion anchors. ‘All bonded anchors require careful attention to cleaning of the drilled holes in order to achieve full load-bearing capacity. In the case of glass capsule anchors, some cleaning of the hole is achieved during installation by treatment of the sides of the hole: with aggregates and fractured glass particles. 3.2.4.1. Capsule anchors. The most common version of the capsule anclior consists of a cylindrical glass ampule containing a polymer resin, an accelerator and mineral aggregate (see Fig. 3.12). Several proprietary methods have been Geveloped to contain the accelerator and resin separately in the capsule, The ccapsule is inserted in a drilled or cored hole; deeper embedments are typically achieved by stacking multiple capsules in the hole. Setting of the anchor is accomplished by direct boring through the capsule with a threaded anchor rod (usually equipped with a chiselled end) chucked directly into 2 rotary rill. Straight reinforcing bars may be installed in the same way. The drilling and hammering action of the drill mixes the contents of the capsule with the fractured fragments of the capsule to form a relatively fastsetting polymer/glass matrix. As well as stiffening the polymer matrix and reducing shrinkage, the fractured glass and aggregate components serve to improve bond by scouting the sides of the hole during installation Capsule-type bonded anchors may’also be set in previously undercut holes. ‘To improve the bond strength, especially in the region of the undercut, the glass capsule may contain short steel fibres. ‘Tube-type systems consist of a plastic or Melinex casing containing two separated and pre-proportioned components which are mixed by kneading of the tube. The tube is placed into the hole after mixing and, as with the ‘capsule anchor, the anchor rodcis: driven into the hole to complete the installation. The end of the anchor rod is usually chisel-pointed to facilitate breakage of the tube. 3.2.4.2. Resin-based injection anchors. Plastic cartridges containing pre- measured amounts of resin and hardener allow controlled mixing of polymer Fig, 3.13. Resin-based components. The components ate typically mixed through a special mixing injection anchor nozzle as they are dispensed, or are completely mixed within the cartridge 2 i Fig, 3.14. Grouted anchors TYPES OF FASTENER immediately before injection (Fig. 3.13). Typically, the catalysed resin is injected into the hole first and the anchor rod (deformed bar or threaded rod) is pushed into the hole and rotated slightly to promote complete contact between rod and mortar. Care must be taken to prevent the formation of air bubbles in the mortar during insertion of the rod, Other systems allow for injection of the mortar after the rod is positioned in the hole. Shims (spiders) ‘can be used for larger anchorages involving longer embedments to preserve the annular space around the rod during injection and setting of the mortar. Other systems utilize a plastic pouch to contain the polymer components, which are mixed by manual kneading of the tube. Immediately after mixing ‘small incision is made in the pouch, and the resin is poured into the hole. Polymer components may also be brought in bulk and mixed either manually ‘or with a power mixer in a bucket and used immediately, or they may be pumped through a mixer and then injected into the hole. Both epoxies and polyester resins are available in bulk packaging, Care must be taken to assure proper mix design and adequate mixing of the resin components on site, and to protect personnel from exposure to fumes and direct contact with the polymer materials. Bolt with nut Bond breaker Threded rod with out 5 Poor wat nt |. Proformed deformed hole 2B INTRODUCTION Fig. 3.15. Plastic anchors with serew as expansion element: (a) examples; (@) installation 3.2.4.3, Grouted anchors, Grouted anchors (see Fig. 3.14) may be headed bolts, threaded rods with a nut and washer at the embedded end (biock-end anchor) or deformed reinforcing bars with or without end anchorage. They ‘are embedded in pre-formed or pre-diilled holes with a Portland cement and sand grout or a commercially available pre-mixed grout. Some proprietary {routs are expansive and thus develop compression stress in the grout after setting, This may serve to increase the contribution of friction to the tension load-bearing resistance of the anchorage. ‘Tf end anchorage is provided, the shank may be either bonded (Fig. 3.14(a)) ‘or debonded (Fig. 3.14(0)). Debonded anchors preclude tension load transfer near the concrete surface, Debonding is typically achieved by means of @ thin sleeve which prevents the resin from contacting the rod. In general, grouted anchors transfer tension loads to the base material by one or a combination of the principal load-transfer mechanisms: bond, friction and ‘mechanical interlock, Belled or keyed holes (Fig. 3.14(¢) and (d)) serve to increase the role of mechanical interlock in resistance to tension loads; however, they can generally be produced only by the placing of blockouts in the concrete formwork, and as such do not technically belong to the post- installed anchorage classification. 3.2.5. Plastic anchors Plastic anchors for fastenings to normal-weight concrete generally comprise ‘an expansion element (usually a screw) and a polymer sleeve slotted at one ‘end to allow expansion. The various forms of plastic anchor available differ in the external form of the sleeve, especially at the expanding end, and in the inner profile of the sleeve. The slotted (expanding) end of the sleeve typically has projecting elements to prevent rotation during installation, The sleeve is expanded by installation of the screw (Fig. 3.15(b)), which forms and cuts a thread into the plastic and presses the sleeve against the wall of the drilled hole, ‘The diameter and thread charateristics of the expansion element are matched to the form of the plastic sleeve. If, in place of the provided expansion element, a normal wood screw is employed, the holding Capacity of the anchor may be reduced significantly. Representative plastic ‘arichors and their corresponding expansion elements are shown in Fig. 3.15(). "The tension toad-bearing performance of plastic anchors depends on correct installation, (a) The temperature of the anchor and the base material must be adequate (minimum 0°C for anchors currently on the market) to prevent brittle failure of the sleeve during expansion. (© The sctows must be inserted completely to develop maximum expansion force. 4 (@) (e) Fig. 3.16, Plastic anchors with nail-like expansion element: (a) examples (©) installation Fig. 3.17. Plastic fasteners ‘for aerated concrete: (a) ‘Plastic anchor with special ‘outer shape and screw as expansion element; ©) plastic anchor with cone as expansion element TYPES OF FASTENER a @ ®) @ CO) (© A complete installation is achieved when neither the screw nor the sleeve will turn with continued application of torque. (@) Once installed, sleeves cannot be reused or reinstalled without significant impairment of their load-carrying capacity. Plastic anchors expanded with a driven expansion element are also available (Fig. 3.16). The installation is relatively quick, but the holding capacity is much less predictable than for anchors expanded with screws. Plastic anchors transfer tension loads by mechanical interlock between the screw and sieeve and by friction between the sleeve and the wall of the drilled hole in the base material Special anchors have been developed for fastenings to aerated concrete (Fig. 3.17@)): they comprise a sleeve with helical ribs and an expansion element (screw). The anchor is driven into a cylindrical drilled hole (drilled ‘without hammering action) and expanded by screwing in the screw. The ribs serve both to prevent rotation of the sleeve during installation and to provide a distribution of stress over a larger area of the base material. ‘The system shown in Fig, 3.17(b) comprises a plug and an expanding device equipped with bolt, nut and washer. By use of a special drill, a cylindrical hole is drilled (without hammering) and then undercut. The anchor is inserted and expanded by tightening of the nut, whereby the expanding device is pressed into the undercut. 3.2.6. Powder-aciuated fasteners Because it requires no electricity, powder-actuated fastening provides flexibility and economy for repetitive applications, Powder-actuated fasteners are nail-like pins or threaded studs (see Fig. 3.18) that ace driven into the concrete or steel by use of a tool powered by an explosive charge. 25 INTRODUCTION Fig. 3.18, Powder-actuated fasteners Fig. 3.19. Operating principles for powder- actuated fasteners: (a) free flight (high velocity); (©) piston (medium and low velocity) 26 i i i ki fT ate i Jnitially, the tools developed to install powder-actuated fasteners functioned like a gun (Fig. 3.19(a)), transferring the cartridge’s energy directly to the pin as kinetic energy. Because free-flying pins endanger the operator and the environment, piston devices have been developed (Fig, 3.19(b)) that transfer the cartridge’s energy to a piston, which in turn drives the pin into the base material. The piston retains 95% of the kinetic energy from the charge ‘and is restrained by the setting device, thus posing less danger to the operator and surroundings. ‘The tools can also be classified as low-velocity, medium-velocity or high- velocity. A low-velocity tool produces a fastener velocity < 100m/s; a Jnediumvelocity tool produces a fastener velocity > 100m/s and < 150sa/s. Jn general, free-flight tools (Fig. 3.19(a)) are high-velocity tools, whereas piston tools (Fig. 3.19(b)) are either low- or medium-velocity. "The operating principle of a powder-actwated fastener is to displace the ‘pase material while penetrating it. In the immediate vicinity of the pin, the structure of the base material will be partially destroyed. The holding power of a powder-actuated fastener in concrete is derived from the compressive reaction forces resulting from the displaced conerete and from chemical ‘bonding between the-conerete and the pin shaft. Due to the high penetration velocity of the pin, the temperature of the concrete at the friction surface is raised significantly; this improves the bonding between the pin and the concrete. TWO aetna Feeney ent sera too om 71 oa | @ Energy required to crive the fastener in Enorgy of piston Energy requied to drive the fastanarin ° TYPES OF FASTENER Fig. 3.20. Powder-actuated ‘fastener anchored in an ‘aggregate While softer aggregates may be penetrated by the powder-actuated fastener (ig. 3.20), harder aggregates and those that are struck by the pin at an unfavourable angle may cause the pin to be deviated anc/or bent (Fig. 3.21(a)). The pin then has a reduced or negligible capacity. Thus, in this case, the pin is generally assumed to have no load-bearing capability (failure during ing). This problem can be avoided if the pins are driven into short pre- drilled holes (depth ~20 mm, hole diameter > shank diameter; see Fig. 3.21(b)). The shafts of the pins must be longer than the depth of the pre-driled hole so that the embedment depth in the base material remains constant. This method precludes failures during setting even in the case of high-strength concrete and/or hard aggregates. Surface damage due to spalling of the concrete is also avoided. In accordance with the greater absolute embedment depths achieved, the ultimate tension loads attained with pre-drilling are generally higher than with conventionally set pins, and the scatter of the ultimate loads is lower. 3.2.7. Concrete screws ‘The concrete screw (see Fig. 3.22) is a threaded anchor available with either a flat or a hex washer head. It varies in length from 32.man to 100mm, and is goncrally available in 4.75 mm and 6.35 mm nominal thread diameters. ‘The threads are case hardened to a hardness of ~ 60 Re, and either are notched or have a high—low thread configuration to ease cutting and reduce resistance uring installation. The anchor cuts threads into concrete or masonry material 4s itis installed. The threads create a mechanical interlock with the concrete or masonry material. Anchor embedment is typically limited to 38mm due to the rapid resistance encountered as the sctew threads cut into the concrete. Fig. 3.21 (lefi). Possible deformation of ponder- ‘actuated fasteners: (a) without predrilling; (b) with predrilling Fig. 3.22. Concrete screw a INTRODUCTION 28 Concrete screw anchors are installed in pilot holes drilled at least 6 mm deeper than the required anchor embedment by use of conerete drill and carbide drill bit. The dust and debris are usually not cleaned out of the holes. Gitedment depth compared with: equation (5.4) (after Bode & Roik (1987)) HEADED ANCHORS IN-UNCRACKED CONCRETE, ‘Standardization Institute (TGL). Some variations, such as the use of different factors (AUS, 1985), may indicate a different interpretation of the predictive capacity of equation (9.3). Use of a pyramidal shape for the failure surface for single anchors (TGL) and anchor groups (PCI and TGL) simplifies the determination of A, in cases involving complex geometries. '5,2.3.2. Method according to Bode & Hanenkamp (1985) and Bode & Roike (1987). Bode & Hanenkamp (1985) and Bode & Roik (1987) present an empirical equation for the average ultimate tension load for single-headed anchors exhibiting concrete-cone failures based on the results of over 100 tests with headed studs. The anchorages tested had an embedment depth range of approximately 40—140mm. Concrete cube strengths f., of 20-45 MPa Were measured at the time of testing. Also recorded at the. time of testing were the corresponding cylinder and splitting tensile strengths fz and fe- Regression analysis of the test data gave the expressions for mean strength 0.96 hes [1 F (dn/hes)] (fee)? (N= (S.4a) Nomieginde = U-B9A? [1 + (dnl) (6) — NY 6.40) Figure 5.11 shows the normalized pull-out strengths N, [fa a8 a function of the anchorage length yyq. Curves for the average pull-out strength Nan according to equation (5.4) are plotted against the data using two boundary values for the ratio of head diameter to embedment depth dy ee “The results of the regression analysis were employed to derive an expression for the characteristic strength (5% fractile, 90% probability) Nexto) = B.DARE [L + (dlls) ee)? (N) (5.5a) Nuxcoytosey = 9-65 AY? TL + Gn/her) 1 (6)°F ®) (6.56) Comparison of equations (5.4) and (5.5) shows the characteristic strength to be ~0.81 times the mean value. Factors to account for the effects of spacing and edge distance are based Nace) ro MONOTONIC LOADING IN CONCRETE * os cs ==: 1 - A | 8 oO i 8 7 | ~ | 8 S 5 Equation 6.8) z oe (7 equation (58) 5 os 3 2 2’ ; V8 = [se ti acoring equation (4) Y : zi 7 BB ee) 10 | | as 08 08 70 os 88 70 eh he Fig. 5.12. Test data for headed anchor groups exhibiting concrete-cone failures compared vith. ‘equation (5.6), plotted as a function of actual-te-critical ‘spacing ratio (ajter Bode & Roik (1987) 46 on an approach originally developed by Bligehausen & Pusill-Wachtsmuth (1982) (Fig. 5.12). Two boundary conditions are assumed for anchor spacing. (@) For a theoretical centre-to-centre anchor spacing of s = 0, the capacity of'a multiple anchor fastening is reduced to the capacity of one anchor. (®) If the anchor spacing s is greater than or equal to a value (4f,)) taken as critical for development of the full concrete-cone failure load, then the group attains its maximum tension capacity, which is n times the failure load of a single anchor. ‘A straight line is assumed between these limiting values, resulting in Mamll + sa— DIG he) 6.6) where n is the number of studs acting together as one group, s is the centre~ to-centre stud spacing and N,,, is the mean failure load of a single anchor according to equation (5.4). Figure 5.12 shows the results of tests with headed stud groups (n = 2, 3, 4 and 9 anchors) loaded in tension through a stiff loading plate (Bode & Roik, 1987). The measured failure loads have been normalized to n times the failure load of a single anchor by use of equation (5.4), and are ploited as a function of the ratio of actual spacing to the assumed critical value of 4hy. The straight line functions for predicted capacity as given by equation (6.6) are plotted against the data in Fig. 9.12. ‘Similarly, a critical edge distance c,, necessary for the development of the full concrete-cone failure load is considered. Two values for the critical edge distance are assumed, depending on the number of proximate free edges. These assumptions give Nasnetge = MerNam Num (6.7) where cy = LShy (one free edge), Cy = 20h (two or more free edges, e.g. anchor in a comer) and Nj.» is the mean failure load of a single anchor (Grom equation (5.4)). Figure 5.13 plots the results of tests conducted with headed studs located near a free edge (Bode & Roik, 1987). The measured failure loads have been normalized using values calculated from equation (5.4), and are plotted in Ne mgrouy Fi 5.13. Test data for headed anchors located near an edge compared with equation (5.7), plotted as a fimetion of the ratio of edge distance to embedment depth (after Bode & Roik f1987)) we HEADED ANCHORS IN UNCRACKED CONCRETE relation to ratio of edge distance to embedment depth c/hy. The bilinear relationship given by equation (5.7) is shown for comparison. 5.2.3.3. Method according to Eligehausen, Fuchs & Mayer (1987/1988) and Rehm et al. (1988) (J method). The ¥ method for predicting the strength of headed anchors exhibiting concrete-cone failures was first proposed by Lehmann (1984). It was revised by Pusill-Wachismuth (1982), Eligehausen & Pusill-Wachtsmuth (1982), and Riemann (1985), and developed to the form presented here by Eligehausen et al, (1987/1988) and Rehm et al. (1988). The method is generally referred to in the source literature as the x method. It has been renamed the y method to conform with CEB notation. | Eligehausen et al. (1987/1988) and Rehm et al. (1988) present @ ‘comprehensive procedure for the calculation of ultimate tension loads for headed anchors exhibiting concrete-cone failures, based on the results of 196 tests with headed studs, The tests involved anchorage depths of 40-525 mm and measured concrete strength f., of 20-60 MPa, Based on a regression ‘analysis ofthe test results, the foliowing relations were obtained for the mean strength Numicate) 1s.5mye fe (NY (5.8a) Nambia = 1TH AEE (5.80) Figure 5.14 shows a histogram of the ratio of measured to predicted failure Joads (Rehim et ai., 1988). The coefficient of variation is given as 14% and the 5% fractile of the failure load is derived as 0.77 times the average value. ‘The equations for characteristic strength (5% fractile, 90% probability) are thus given as Nexon = 12hH fe? ON) 6.98) Nexioinsey = 13hG? F25 ON) (6.96) ‘As with the method described in ACI 349, equation (5.8) assumes the failure Joad to depend solely on the concrete tensile strength (taken as proportional to f9%) and the embedment depth. i in Fig, 5.15 (Rehm et al., 1988), test results for headed studs exhibiting concrete-cone failures are plotted as a function of embedment depth hi. The ‘curve for mean ultimate strength given by equation (5 8a) is plotted against the data for comparison. The tests were carried out in concretes of varying strengths; the measured failure loads have been normalized to a concrete strength fi, = 25MPa using the square root of the compression strength 08; NtesiNam os Equation (5.7) Namacoording to equation (5.4) 7s het ” 1200 ~ 4 ke 225A na = NA2sit 1000} - 2001 — —f~$— ‘according 10 equation (5.8) for headed anchors exhibiting concrete-cone failures: test results from Fig. 5.14. Ratio of actual to predicted tension capacity ag Single anchors without edge influences (afier Rehm ob et al., 1988) Fig. 5.15 {above right). Comparison of equation (6.8) with tension test results for headed anchors exhibiting concrete-cone failures normalized t0 a ‘concrete strength f.. = 25MPa and plotied as a function of embedment depth (after Rehm et ala, 1988) 48 Nyt = 18h! See°9/1000 | 600 |— 400) - aoe tA 200 a0 00 ratio. As such, Fig. 5.15 does not indicate the accuracy of the assumed relationship between failure load and concrete strength (N, & Vf.) inherent in equation (5.8), but it does indicate the dependency of N, on hy? as derived from the regression analysis. ‘AS with the method of Bode & Hanenkamp (1985), allowance for the effects of edge distance and spacing is based on the approach originally developed by Eligehausen & Pusill-Wachtsmuth (1982). Based on an assumed crack angle for the failure cone of « ~ 55°, and supposing that overlap of adjacent failure cones leads to a reduction in group capacity, the investigators proposed the critical spacing s,, = 3h (Cf. sa, = 4h, (Bode & Roik, 1983)) to avoid any spacing influence. A straight line relationship is assumed between the boundary conditions of s., = 3 and s., = 0, which results in the following ‘equation for the mean strength of a ‘group’ of two anchors Nasmgrup = Vso (5.10) wo Dt Shey 52 6.1 where Jay = Shy and N,.. is the mean failure load of a single anchor from equation (5.8). Figure 5.16 (Rehm ef al., 1988) plots test results for headed anchor groups (two anchors) loaded in tension through a stiff loading plate as a function of anchor spacing divided by embedment depth s/hye. All tests ended in cconcrete-cone failure. The ultimate loads obtained from the tests have been divided by the calculated failure load for a single anchor by use of equation (5.8) to normalize the results with respect to concrete strength and to facilitate comparison with the bilinear relationship given by equation (5.10). For the calculation of the ultimate load of orthogonally arranged quadruple fastenings (see Fig. 5.17), Rehm et al. (1988) proposed that factors yj, be separately derived for each orthogonal direction and multiplied as follows Nemaiop = Wai ¥aNom 6.12) | Fig. 5.16. Comparison of equation (5.10) with test results for headed stud groups (two anchors) loaded in tension through a stiff loading plate: test results have been divided by the calculated faiture load for a Single anchor using ‘equation (5.8) and plovted (a5 a function of anchor spacing divided by ‘mbedment depth (after Rehm et al., 1988) © Fig, 5.17, Concrete-cone failure of a quadruple ‘fastening HEADED ANCHORS IN,UNCRACKED CONCRETE Va = 1+ 5h5q) $2 (8.13) where 5; is the centre-to-centre spacing measured in direction i and Nm is the mean failure load of a single anchor from equation (5.8). Figure 5.18 (Rehm et al., 1988) plots the tension failure loads from tests on quadruple fastenings with equal spacing in two directions as a function of anchor spacing divided by embedment depth. Tests with headed, expansion and undercut anchors are included. As with Fig. 5.16, the data have been normalized with the failure load for a single fastening calculated from equation (6.8). Equation (5.12) data are plotted against the test data for comparison. Implicit in equation (5.12) is the assumption that negative effects due to spacing (ie. overlapping of stress fields) and other factors (e.g. edge distance) on load capacity can be accounted for by multiplication of individual effects determined through measurement or analysis.’ This assumption of a multiplicative effect forms the basis for the ¥ method, Rehm ef al. (1988) extend this idea to orthogonally and symmetrically arranged multiple fastenings with a maximum spacing in each direction s,, < sq by substituting the distance between the outermost anchors of the group 5, for s; in equation (6.13) and maltiplying all terms as follows None = Th VesNe 6.14) where Nag is the average failure load of a multiple anchor group with anchors, Nm is the mean failure load of a single anchor from equation (5.8) 8 T T Num according to equation (6.8) el ° | jt 3 & 3 Equation (6.1 2 :quation (6.10) 1 oy | ait ©: Anchor 1 Headed etud 1 2 3 thot Penn eee eee MONOTONIC LOADING IN CONCRETE Fig. 5.18. Comparison of ‘equation (5.12) with test results for quadruple anchor 3 ‘groups loaded in tension 4 - through a stiff loading 7 plate: test results have been ‘Nugn aocording to onustion (5.8) e divided by the calculated ° Jailure load for a single anchor using equation (5.8) ‘and plotted as a function of ‘anchor spacing divided by embedment depth (after Rehm et al., 1988) C : N&: ’ 3 a) [ese] eunton 6.12) Lo Anchor . 1B Headed stud 4 sitet predicted failure load for groups of headed anchors &s.a fsction of the spacing ofthe outermost anchors ° § (afer Eligehausen et al, e, 19%) Fig. 5.19. Ratio of actual 10 E bee. aroun Meal grour o Fauds 2 see 2 testucs © 36 studs fiat 185 mm ‘Mxcakgroup aocording to equation (5. sane S28 Fig. 5.20. Model for the influence of load ‘eccentricity on the failure load for a group of two infinitely stiff anchors loaded through a rigid plate (according to Riemann, 1985) Nya WeDo Nye = Woven My2= Nam Went + alse 52 Vo= tt + 2etse) 50 HEADED ANCHORS IN UNCRACKED CONCRETE Bog = AF SiulSee S Mi 6.15) where s,, is the distance between outermost anchors in direction i (see Fig. 5.19), n-is the number of anchors in direction i, m = 1 for anchors in one tow and m, = 2 for anchors in two directions. As before, all anchors in the group are assumed to be loaded equally. Figure 5.19 (Eligchausen et al., 1992) plots the ratio of actual to predicted tension capacity of groups of headed anchors in relation to the spacing of the outermost anchors s,. Group fastenings with up to 36 anchors were loaded concentrically in tension through a rigid loading plate to assure equal oad distribution. The embedment depth A was held constant at 185 mm. ‘The spacing of the outer anchors was 100-875 mm; the spacing of the individual anchors was 100-400mm (0.54h.-—2.2he). The concrete strength f, was ~25MPa. All anchorages exhibited concrete-cone failures. Figure 5.20 shows the influence of load eccentricity on the carrying behaviour of anchor groups. With centric loading (case 1), equation (5.14) is valid. When the load acts on one headed stud (case IN), the limit load of the group corresponds to the value for an individual anchor, irrespective of the axial distance. For a tensile force acting anywhere within the group (case I), a hyperbolic curve between the limiting cases I and IIT can be assumed for the failure load of the anchor group. The influence of eccentricity ‘on the rupture load of the group can thus be determined by use of an additional coefficient Y. Numa = ¥e¥eNem 6.16) where N,,., is the mean failure load of a single anchor from equation (5.8) ve = WEL + Gels) = 1 6.17) where e is the eccentricity of tension force from the centroid of anchors (see Fig. 5.20); ¢ < 5/2. Equation (5.16) can also be used for groups of any prevent’splitting. faire of an anchorage in a reinforced concrete member. ‘While the work of Hasselwander et al. (1987) included tests with anchors in specially. reinforced specimens, the testing was limited and no attempt was made to isolate the influence of the reinforcement on the anchorage: strength. Although an equation was presented for the strength of the attchoragés'tested, it must be assumed that this equation is valid only for the reinforcing arrangements tested. Rehini et al: (1988) have proposed that splitting failures can be avoided in anreinforced members with concrete strength f. greater than 20 MPa.if edge distance > 1.5hy, spacing = 3.0hy and member thickness = 2.0 hey. 5.3.1. Types of failure and load—displacement behaviour ‘The failure’ modes of headed anchors loaded in shear can be categorized as follows (see Fig. 5.37) 6 MONOTONIC LOADING IN CONCRETE Fig. 5.37, Failure modes of headed anchors loaded in shear Fig. 5.38. Shelt-shaped spall and associated shaft deformation of a headed stud subjected to shear loading (from Fuchs, 1990) a 5.3.1.1. Steel failure (Fig. 5.37(a)). For deeper embedments, failure is generally characterized by bending of the anchor shaft. Due to the locally high pressure in front of the bolt, a conchiform concrete spall may occur near the upper surface of the concrete before maximum load is obtained. With increasing load, the anchor shaft yields and/or fractures. Depending ‘on the concrete strength and the size, embedment and steel qualities of the anchor, deformations at failure may be relatively large. The failure load increases with increasing steel strength and cross-sectional area. 5.3.1.2. Concrete failure (Fig. 5.37(6)— (g)). Failure is characterized by one of three possible modes. (@) Concrete spalting is shown in Fig, 5.37(a) (see also Fig. 5.38). Various authors define the onset of concrete spalling as a point of failure for the anchorage based on serviceability considerations. Particularly for deeper anchorages, the maximum load may be reached at much greater displacements and. anchor deformations. (©) Lateral.concrete-cone failures due to edge proximity are shown in Fig. 5.370). ‘The original of the lateral cone may be at the head of the 1 i j Fig. 5.39. Typical load— displacement curves of headed anchors under tension and shear loading (after Hawkins (1987)) HEADED ANCHORS IN: UNCRACKED CONCRETE Loa kN + shear (ky): ste0t fare —O— Shear {kt concrete fale Displacement. mm anchor for short stiff embedments; for deeper embedments it usually begins closer to the surface of the concrete. If the development of the full concrete cone is limited by spacing or edge condition (Fig. 5.37(c)—@)) ot by the depth of the member (Fig. 5.37(f), the failure load is generally reduced. As with tension loading, the failure load is influenced by the concrete tensile capacity. Other influencing factors include the side cover c, the flexural stiffness of the anchor shaft and the embedment depth. (© Concrete crushing combined with fracture of the concrete behind the anchor and Subsequent pull-out of the anchor is shown in Fig. 5.37(g). ‘This failure mode generally occurs with simall embedment depths (fer 4d—6d) or anchor groups with small spacing. 5.3.1.3. The typical load—displacement bchavionr of a single headed stud loaded in ‘pure’ shear can be described as follows (see Fig. 5.39). On initial loading, the shaft of the anchor beats directly on the surface concrete, this being the stiffest load path. As the surface concrete is crushed, flexural stresses are generated in the anchor shaft as it attempts to redistribute the load over aa greater depth. Increasing load will usually lead to the formation of a shell- shaped spall (see Fig. 5.38), the depth of which is determined primarily by the flexural stiffness of the anchor shaft. Spalling increases the flexural demand on the anchor shaft and simultaneously reduces the internal level arm, For shallow anchors, the onset of spalling may be accompanied by a flexural/shear failure of the anchor shaft or concrete fracture originating at the anchor head similar to that shown in Fig. 5.37(g). In the case of headed anchors with sufficiently large embedment depth and shaft strength, however, the load may continue to increase until the anchor is sheared off through a combination of shear, flexural and tensile stresses. ‘The displacements associated with this post-spalling behaviour may be quite large, and in general the mode of shear load transfer (concrete crushing and spalling followed by anchor shaft bending) leads to larger displacements at failure than those associated with tension loading. In the case of a headed anchor loaded in the direction of a free edge and exhibiting « lateral concrete-cone failure, the ultimate load (and corresponding displacement) increases with increasing edge distance. For a given anchor, however, the load associated with the failure mode described in section 5.3.1.1 (palling followed by steel failure) represents an upper bound on the anchorage shear capacity 6 | MONOTONIC LOADING IN CONCRETE Table 5.1. Shear equations and values of the shear reduction factor 5.3.2. Failure loads — steel failure Headed anchors subjected to shear loads experience shear, bending and axial stresses. There is no generally accepted theoretical formulation for the calculation of the rupture load. It is often assumed that failure occurs when the bending stresses exceed the steel tensile strength: the bending moment has been calculated for the case of a beam on an elastic or elasto-plastic support Gasler & Witta, 1967; Cziesielski & Friedmann, 1983; Friberg, 1940; Wiedenroth, 1971). The rupture loads thus determined, however, rarely match the test results (Fuchs, 1984). It seems, therefore, that empirical formulations are more useful at present. ‘The literature presents several methods that consider the failure load solely 1s a function of the steel strength, and one method that includes the concept of shear—friction. These approaches are reviewed in sections 5.3.2.1 and 53.2.2. ‘5.3.2.1. Method according to ACI 349-85 (1985) (shear—friction theory). ‘The shear capacity for embedments is calculated from Ve= AS, (bf, N) 6.31) where ¢ is a strength reduction factor (@ = 0.55 for structural shapes, fabricated steel sections and shear lugs embedded in the concrete, and = (0.85 for bolts, studs and bars), A, is the total cross-sectional area of headed anchors attached to embedments, f, is the steel yield strength (= 120 kgffin’ (~828MPa)) and y. is the coefficient of friction. The term ‘embedment’ as used by ACI 349-85 refers to an anchor plate with anchors either welded or bolted to the plate. The coefficient of friction is typically 4 = 0.9 for concrete or grout against as-rolled steel with the contact plane a full baseplate thickness below the concrete surface, « = 0.7 for concrete or group against as-rolled steel with the contact plane coincidental with the concrete surface, and = 0.55 for grouted conditions with the contact plane between grout and as-rolled steel exterior to the concrete surface. Tf the friction coefficient is taken as 0.7 for steel against concrete, the ultimate shear strength ( = 1.0) of the embedment is given by Veo = 0.7 Arf, (bf, N) 6.32) ‘5.3.2.2; Methods not based'on shear—friction theory. Many investigators have found that the average steel failure load of a headed anchor loaded in shear can be predicted by Visa Af, (bf, N) (5.33) ‘Various forms of this equation in the literature relate the sheat reduction factor ato either the ultimate or the yield strength of the steel. Table 5.1 gives Reference ‘Test basis | ‘Shear equation « Burdette et al. (1987) Steelleonerete Van = adh 0.65 ‘Chesson, Faustino & Munse (1965) Steel/steel OAS, 0.55—-0.64 ‘Fuchs (1984) ‘Steel/concrete i OAS, 06 Hencky, Huber & Mises Theory Vin = adh, 0.58 ADT (1985) Steel/concrete Vom = OAs, 0.7 Kiingnet, Mendonca & Malik (1982) Steel/conerete Vins adift 0.675 Kulak, Fisher & Struik (1987) ‘Steel/stee! Vin = OAS, 0.62 McMackin, Slutter & Fisher (1973) Steel/conerete Vin = oh 10 Roik (1982) Steel/concrete Ven = oA, 07 ‘Shoup er al, (1963) ‘Stee/conerete Vom = OAS, 1.0 TVA (1975) ‘Steel/grout/concrete Vin = Af, 0.53 Valtinat (1982) Screws Jom = OAS, 0.625 6 Fig. 5.40, Comparison of ‘actual failure loads for anchors loaded in shear towards a free edge and values predicted from equation (5.34) (after Klingner & Mendonca, 19826) HEADED ANCHORS IN-UNCRACKED CONCRETE these equations with the cortesponding shear reduction factors proposed by the authors. "The calculated strength of a group of headed anchors exhibiting steel failure depends on the distribution of load to the anchors. In general, when equal load distribution is assured (as when studs are welded to a concentrically- loaded anchor plate), the failure load is » times the value obtained from equation (5.28) (n = number of fastening elements). Fuchs (1990) investigated the case of two anchors projecting through an anchor plate loaded in shear with variable clearances between the holes in the anchor plate, and the anchor shafts. Expansion anchors and undercut anchors manufactured from high-strength brittle steels were employed in the tests, which showed reductions in group strength of up (0 20% for hole clearance differences of up to 4mm. The degree of reduction is related to the ability of the anchors to redistribute load through plastic deformation of the anchor shafts. 7 Cook & Klingner (1989) propose that the shear strength in a multiple-anchor connection should be taken as 50% of the tensile strength (V,/T, = 0.50). ‘The shear strength (steel failure) of undercut anchors in a multiple-anchor connection should be taken as 60% of the tensile strength (V,/T, = 0.60). 5.3.3. Failure loads — lateral concrete-cone failure (edge failure) In the case of headed anchors loaded towards a free edge, a lateral cone failure may pre-empt failure of the steel (see Fig. 5.40). According to Fuchs & Eligehausen (1986a) and Stichting Bouwresearch (1971), the angle of fracture as measured from a line passing through the anchor and parallel to the edge is approximately 30°—35? on average, and the depth of the lateral failure cone is approximately 1.3—1.5 times the edge distance. Klingner & Mendonca (1982b) report the angle as ~ 45° for edge distances of 125mm or more, decreasing linearly to 25° as the edge distance approaches zero. ACI 349-85 (1985) assumes an angle of 45° ‘The ultimate load capacity associated with this failure mode depends on the behaviour of concrete under multiaxial stress. As in the case of tension loading, a complete analytical determination of the concrete failure load is not yet available. The following equations arc therefore derived empirically, taking account of observed behaviour. As in the case of tension loading, these fracture angle measurements are the basis for the spacing requirements associated With the various methods described in sections 5.3.3.1-5.3.3.4. ‘The methods are compared in sections 5.3.3.5—5.3.3.7. Other factors are discussed in sections 5.3.3.8 and 5.3.3.9. 5.3.3.1. Method according to ACI 349-85 (1985). The average strength of an anchor exhibiting @ lateral concrete-cone failure is calculated by assuming a uniform stress of 4(/2)"* Ibffin’ on the projected area of the failure cone Vom = G40) Ae (bf 6.34) (Hom = 90.1650.) 40) ) aor Specimens with some pravious damage 10 Number of esis 2 3 Aciualipredicted capacity 6s ere eee MONOTONIC LOADING IN CONCRETE Fig. 5.41. Comparison of ‘actual failure loads for deformed reinforcing bars loaded in shear towards a ree edge, and values ‘predicted from equation (65.36) (after Paschen & Schonhoff, 1983): n = 38, x= 1.06, 9 = 14% where @ is the strength reduction factor (= 0.85), fis the specified cylinder compressive strength of concrete (Ibf/in*), Ae is the projected area (in?) of the 45° lateral failure cone(s) radiating towards the free edge surface from the bearing surface of the anchor shaft(s), reduced for overlapping areas (i.c. ‘multiple anchors) and intersections of the cone(s) with concrete surfaces (see section 5.2.3.1 for a similar determination of the failure cone for tension loading) and A, = 7c (c + d,)/2 for a single anchor of shaft diameter d, located a distance c from the free edge. When ¢ is taken as 1.0, equation (5.34) is intended to predict the average failure load Where spacing and edge distances are not sufficient to allow the full development of the assumed failure surface, a reduction in capacity directly proportional to the reduction in surface area is assumed. The ACI approach assumes a failure angle of 45°, and the maximum capacity is achieved for a member thickness A > c, Anchors in a row parallel to the edge are assumed to have no influence on each other if the spacing is sufficient to prevent overlap of the failure cones (s., = 2c). This assumption also leads to the conclusion that a single anchor is not affected by a comer if the edge distance c perpendicular to the load direction is at least equal to the edge distance ¢ in the load direction. ‘According to Klingner & Mendonca (1982b), equation (5.34) is valid only for ‘fully embedded’ anchors, full embedment being defined as the embedment that will yield a steel failure for tension loading. Klingner & Mendonca (1982b) ‘compared actual and predicted strengths for tests with headed anchors loaded in shear, using equation (5.34) (see Fig. 9.40). 5.3.3.2, Method according to Shaik & Whayong (1985). Based on analyses ‘of test results from MeMackin et al. (1973) and Klingner & Mendonca (1982b) swith headed studs, Shaik & Whayong (1985) propose the empirical equation Vom = 125 FDS cl — db) (6.35) (am = 5:25 (f% e) ON) In the data considered, the concrete strength was 40505270 Ibffin* (£28MPa—~36MPa) and the anchor diameter was 0.75—2in (~19 mm—~51 mm). 5.3.3.3, Method according to Paschen & Schonhoff (1983). The following emipitical equations are proposed, based on analyses of results from tests with straight deformed reinforcing bars (d, = 14—40ram) in concrete of strength fe = 25-45 MPa: for ple < 1.73 Vim = (190 + 0.230) (f,)°% sin 0.91 cx/c) CN) (5.362) Va = (212 + 0.26c7)()° sin (0.91 cfc) (N) (5.36b) AZ WIA. eo 10 12 13 1 valved casas Fig. 5.42. Test results for various anchor types loaded in shear towards a free edge, normalized to d = I8mm, hy = 80mm and f = 2014Pa by use of the terms from equation (5.38) (afer Eligehausen & Fucks, 1988) HEADED ANCHORS IN UNCRACKED CONCRETE and for cle, 2 1.73 Bam = (190 + 0.23¢4)(f..)°" ™ (8.378) = 212 + 0.264) 9 m (5.37) where ¢, is the distance from anchor to free edge parallel to shear load and ‘ys the distance from anchor to free edge perpendicular to shear load. Fig. 5.41 (Paschen & Schénhoff, 1983) shows a histogram of actual~predicted capacity ratios for equation (5.36). 5.3.3.4. y method according to Bligehausen & Fuchs (1988) and Fuchs (1990). (This method is generally referred to in the source literature as the -« method. It has been renamed the y/ method to conform with CEB notation.) ‘Based on regression analyses of 147 tests with headed, expansion and bonded anchors, Eligehausen & Fuchs (1988) and Fuchs (1990) propose the following ‘equation for the calculation of ayerage ultimate failure load of a single anchor loaded in shear towards the edge Vas = d*F2 (ld )9%Ch (6.380) Vim = LL df25 (Mgeid)?7e!> (.38b) ‘The database contains details of anchors of diameter 8—50 mm and of concrete strength approximately 10-50 MPa. The authors propose that this equation is valid for embedment depths 4d < hr < 8d and member thickness h = 14d. For anchors of embedment length hg ~ 4d, equation (5.38a) simplifies to Vom = 13d p5c'4 (6.392) Vem = 14d pS 6.390) Figure 5.42 (Bligehausen & Fuchs, 1988) compares the failure loads predicted by equation (5.38) and the database used for the. regression analysis. The thickness of all test specimens was greater than 1.4c; The test results were normalized to a concrete strength f, = 20MPa, a diameter d = 18 mm and an embedment depth hy = 80mm by use of factors (20/f,)°5, (18/d)°$ and (has/d)°?. Fig. 5.42 is therefore valid only for judging the accuracy of the term c' relative to the database. ‘The y method (see section 5.2.3.3) has been extended by Zhao, Fuchs & Hligehausen (1989) to shear loading by the introduction of a factor to reduce the failure load for member thicknesses less than 1.4¢ 200 © Non-headed anchor Equation (5.98) fa 9 Headed stud s 2 100 ~ 3 ° 2 | I . tj i i | "fe mH 708 300 300 a a MONOTONIC LOADING IN CONCRETE Fig. 5.43. Test results for 1 expansion anchors loaded in Shear tonards a fee edge 1s0/- v nm and exhibiting lazeral rolage concrete-cone failure, of | = plotted for various member 73 thicknesses as a function of gal (og ‘edge distance (after Zhao et al, 1989) dae \e “+ a8 “| ae 700 200 300 00 Fig. 5.44, Comparison of ‘equation (5.41) with test rresulis for expansion anchors loaded in shear fa 1 towards a free edge lo 8 | ‘exhibiting lateral concrete- a Pew] ‘cone failure: rest results have been normalized by the °° a] o 3 value Vig predicted by Equation (541) ‘equation (5.38) and are plowed as a function of hc ~ a4 (after Zhao et al., 1989). ] Via is the average failure lead for a large member dept h = 1-40) o| L 0 rs 35 32 is 20 ne ValVonn Fig. 5.45. Comparison of equation (5.42) with test hhave been normalized by the results for expansion anchor 2g (groups loaded in shear towards a free edge exhibiting lateral concrete- 8 [ uaton 42) ew vale Vay predicted by 1 equation (5.38), and are Vageore!*Mam ‘cone failure: test results v plotted as a function of sic ~~ = (after Eligehausen & Fuchs, - fe 1988) ° t 2 a Fig. 5.46. Anchors in a row parallel to an edge loaded eccentrically (see equation (5.44)) (after Bligehausen & Fucks, 1988) 6 Fig. 5.47, Failure mode of «a single anchor loaded in shear when located in. (a) a corner; (b) a narrow member HEADED ANCHORS IN .UNCRACKED CONCRETE Vasa = VV (5.40) Ve TL 4e)°" = 1.0 6.41) where V,. is the average failure load for a single anchor loaded in shear, from equation (5.38). ‘The data used to develop this relationship come from tests with torque- controlled expansion anchors. Zhao et al. (1989) assume that the results are equally valid for headed anchors. Fig. 5.43 plots test results against curves for various values of h resulting from equation (5.40); Fig. 5.44 shows the same test results normalized by the average failure load with h = 1.4c from equation (5.38) arid plotted as a function of h/c, The curve given by equation (6.41) is plotted for comparison. Eligehausen & Fuchs (1988) have proposed a further factor ¥, to account for anchor spacing. Similarly to the approach proposed for headed anchors in tension, a critical value s,,*is assumed to develop the full strength of anchors exhibiting lateral concrete-cone failure. This leads to the following relation for two anchors loaded in shear towards a free edge Vagroup = WsKoon (5.42) Ve 1+ sls, $2 (5.43) where s = 3¢ and Vig is the average failure load for a single anchor Toaded in shear, from equation (5.38). Figure 5.45 (Bligehausen & Fuchs, 1988) gives test data for groups of two anchors loaded toward a free edge normalized with the predicted load for a single anchor from equation (5.38) and plotted as @ function of s/c. The strength predicted by equation (5.42) is plotted for cemparison. This approach is extended to a larger number of anchors with an eccentrically applied load (see Fig. 5.46) by substituting the spacing of the outermost anchors for and introducing the term y, (see séction 5.2.3.3) as follows:(Eligehausen & Fuchs, 1988; Fuchs, 1990) Vagioap = Vee Varn (6.44) % = ssa sn 6.45) Ye = UG + Deh) : 6.46) where sq = 3c, 5; is the spacing of cutermost anchors’ (individual anchor spacing § < 3c), ¢ is the eccentricity of load with respect to the centre of gravity of the anchor group (e < s,/2), n is the number of anchors ina row parallel to the loaded concrete edge, and V,,q is the average failuite load for a single anchor loaded in shear, from equation (5.38). No experimental data for larger anchor groups with/without eccentric loading are available to verify the accuracy of equation (5.44). For the case of anchors located near a corner and loaded in shear towards one edge (see Fig. 5.47(a)), Bligehausen & Fuchs (1988) predict a reduction in the failure load given by equation (5.38) if the distance to the adjacent edge is less than 1.5c. The boundary case of an edge distance c, = 0 is ee @ © MONOTONIC LOADING IN CONCRETE : 3 7 ' i se Swati 2 7 8 a Siete i =} et etcas ies : f typ dos 8 Hs yk, aap ol 7 ol i ° 05 1 18 ° Os 76 1 20 25 calor cxier Fig. 5.48 (above left). Test assumed to result in a reduction of the failure load by 70%. This leads to resuls for expansion anchors located in a corner Vessornee = WeVam 6.47) ‘and loaded in shear (Fuchs, = el) & Recah aa cane hans ts 0.3 + O.Feqleq) $1 (6.48) been normalized with the where cq = 1.5¢, and V9 is the average failure load for a single anchor value of ¥.,from equation joaded in shear, from equation (5.38). Limited data from tests with expansion (5.38). (after Fuchs, 1990) anchors located in a corer (Fuchs, 1990) are presented in Fig. 5.48. The experimental failure loads are normalized using equation (5.38) and plotted Hig. 5.49 (above righ) as a function of c;/cj. For the tested case c, = cy, equation (5.47) gives @ gi ‘anmparison of fst results et Cameron Schonnosy Yasser = 0.17 Var f Po ictemmacion (330): A further case to be considered with respect to the ¥ method is that of eee hese on deformed a anchor located in a narrow member, as shown in Fig. 5.4700). The reinforcing bars loaded in expression provided for this case. (Rehm et al,, 1988) is Shear ina narrow member: the ulimate loads have been Vasoan enter = ¥eVase 6.49) normalized with the value of yy teeta 5.50) Van from equation (5.38) where ce, = 1.5¢, and V4 is the average failure load for a single anchor loaded in shear, from equation (5.38), with ¢ = ¢). Figure 5.49 plots the results of tests by Paschen & Schénhoff (1983) on deformed reinforcing bars in narrow members. The data have been normalized by the value for V,,_ resulting from equation (5.38) and plotted as a function of cy/cy. In Fig. 5.49, the available data consider only the case of equal edge distances. In the case of unequal edge distances (C2, # ¢,2) where the Gifference is relatively small, Eligehausen & Fuchs (1988) propose that the average values of c,,, and c,2 inserted in equation (5.49) will yield a conservative result for the ultimate load. Fuchs'(1990) cites a lack of test results for anchor groups located in comers Fig. 5.50. Failure pattern a for an anchor group loaded fata in shear towards a free mea? edge (from Rekm et al., 1988) he Fracture crack 2 Row 2—= 0-——-0-===0 Fracture erack 1 Row ie ee 0 HEADED ANCHORS IN UNCRACKED CONCRETE, Yaron Load Yayoup: kN f= 25 MPa Hote clearance =2 mm 2 Fig. 5.51. Typical load— displacement curve for an anchorage comprising two anchors aligned perpendicular to a free edge ‘and loaded in shear (after Fuchs & Eligehausen, 1988) Fig. 5.52. Predictions for failure load of an anchor loaded in shear exhibiting lateral concrete-cone failure: f, = 25 MPa, d= Tamm, hie > LA 2 4 Displacement: mm and narrow members, but proposes that the multiplicative effect of the ¥ method will also yield conservative results for this case. As discussed above, a concentric shear load applied to anchors such as headed studs welded to an anchor plate will be distributed approximately equally to all anchors. If the anchor group is loaded in shear towards a free edge, the anchors nearest the edge will fail first (soe Fig. 5.50). The corresponding load carried by the anchorage at this point is given by Rehm et al. (1988) as Va.eroop = Mf + 5f(3.00%,1)1 Vo, (5.51) where ny is the number of anchors arranged one behind another in the direction of the load V. Following failure of the anchors closest to the edge, the shear force is cartied by the more remote anchors, and the maximum load is reached when a parallel lateral failure cone is formed (crack 2 in Fig. 5.50). Eligehausen & Fuchs (988) report that, given typical spacings (¢1,1,¢i,2), ctack 1 has only a slight influence on the rupture load of the group. This implies that the concrete failure load for the group depends only on the row of anchors further from the edge, and can be calculated by the use of equation (5.42), taking ¢ = C1, 28 the distance from the edge. In the case of anchorages in which the anchors are not welded, but project through holes in the anchor plate, some construction clearance is typically provided between the anchors and the plate. In this instance the load—slip behaviour of the anchorage is reported by Eligehausen & Fuchs (1988) to depend significantly on which anchors are initially in contact with the anchor plate, For the simplest case of two anchors aligned perpendicular to the edge with only the anchor nearer the edge initially in contact with the loading plate, the observed load—displacement behaviour of the group is shown in Fig. 4007 a — actato.as (1085) a= Shates wnayong (1988) a= Pasehon a Senannot (1983) o= erganausen Fucts (1888) i 4 3 -200- le 5 400- 180 ‘200 sto Distance to fre edge & mm “ato n MONOTONIC LOADING IN CONCRETE Fig. 5.53. Predictions for 19) - ultimate strength of an ee ‘anchor located ina comer BA exhibiting lateral concrete L { Gone failure C4 | Y 7 Y er | g a 2 E os 4 , : 7 2 7 / ; {AC 389-05 (1995) / —=— Eligehausen & Fuchs (1988) TS Pasenen & Schon (963) °% 1 2 fer Fig. 5.54, Predictions for timate strength of @ group of two anchors exibiing lateral concrete-cone failure = t = Act 6#0-85 1808) = === Eigenauson & Fuchs (198) 1 °9 1 2 3 wc Fig. 5.55. influence of a 10 reemant on the load— displacement Behaviour of headed anchors loaded in shear (from Rehm et al., 1988): fe, ~ 281APa, dy = Tamm, d = 22mm 1 ® S od ote Fa ® % 5 10 1s Displacement: mm n HEADED ANCHORS IN UNCRACKED CONCRETE 5.51 (Bligehausen & Fuchs, 1988). Typically, failure of the conerete in front Of the anchor closer to the edge is followed by load transfer to the second anchor. Since the load is again carried by the anchors in series, the maximum load as given by the second peak in Fig. 5.51 is determined by the anchor further from the edge. Since the anchor further from the edge determines the maximum load, this is theoretically independent of the order in which the anchors come in contact with the plate during loading, and equation (5.42) is still applicable. 5.3.3.5. Comparison of different approaches. Figure 5.52 compares the average ultimate shear loads for single-headed anchors exhibiting lateral “concrete-cone failure predicted by the methods described in sections '5.3.3.1—5.3.3.4, For the usual concrete strength (C25) assumed for this ‘comparison, the equations proposed by Shaik & Whayong (1985) and Fligehausen & Fuchs (1988) are almost identical. According to ACI 349-85 (1985) and Paschen & Schémhoff (1983), the shear capacity of an anchor exhibiting lateral cone concrete failure increases proportionally to c*. This leads to higher failure Joad values than the predictions of other methods for edge distances = 100mm. Except for that of Paschen & Schénhoff (1983), all the methods assume that the failure load is directly proportional to the square root of the ‘compression strength of the conerete. Paschen & Schénhoff (1983) assume proportionality of the failure load to the term f2°. The tests on which they based their conclusions were conducted in concrete with litte strength variation, ‘The predictions for failure load given by the various methods for the case of a single anchor in a corner are compared in Fig. 5.53. The calculated values are divided by the failure load assuming no edge influence, and are plotted as a function of c2/c,. The methods of Paschen & Schénhoff (1983) and Bligehausen & Fuchs (1988) agree quite well for the range (c2/c\) > 0.5. By virtue of the model used to calculate the reduced load, ACI 349-85 (1985) assumes a much smaller influence ‘of edges on the failure load than the other methods. Figure 5.54 compares the predictions of the various methods for anchor groups (two anchors) exhibiting lateral concrete-cone failure. The reduced failure loads are divided by the failure loads for two single anchors without spacing influence as predicted by the proposals, and are plotted as a function ‘of s/c. As with the case of an anchor located in a corner, the ACI 349-85 (1985) equation predicts higher capacity than the relation given by Eligehauscn & Fuchs (1988) over the range (s/c) <3. 5.3.3.6. Reinforcement, The sheat load-bearing capacity of headed anchors located near the edge of a member can be increased by appropriate reinforcement, This is shown in Fig. 3.55 (Rehm et al., 1988), which plots typical load—displacement citrves for headed anchors (shank diameter d = 22mm) with a distance from the edge of ~75mm. While the increased capacity achieved by the reinforced anchorage represented by curve (2) is significant, it is far less than would be expected if complete load transfer to the stirrups on either side of the anchorage had been achieved. The investigators concluded that this is explained by the poor anchorage of the stirrups in the area of the lateral failure cone. Curve (3) demonstrates the significant load increase given by the introduction of hairpin reinforcement around the anchorage. In this case, however, the ultimate load attained was still below the calculated capacity of the hairpin. The investigators teported that subsequent examination of the concrete between the anchor and the hairpin revealed that it had been crushed. Paschen & Schénhoff (1983) conducted similar tests in which the hairpin was placed in direct contact with the anchorage. It was found that the yield strength of the hairpin could be developed given adequate hairpin enchorage a MONOTONIC LOADING IN CONCRETE Fig. 5.56. Recommended reinforcement for anchors near an edge loaded in shear (from Paschen & Schonhoff, 1983): ty is the development lengtit ‘according to CEB-FIP Model Code (1990) ™ hs teem em v vot a ¢ | nad 50 mn | beyond the zone of failed concrete (see Fig, 5.56). Klingner et al. (1982) also conducted tests on anchors in direct contact with hairpin reinforcement They found the proximity of the hairpin to the top surface of the concrete to be a significant factor in the performance of the anchorage. Anchorages with the hairpin placed within 20 mm of the surface showed a far greater increase in both stiffness and strength than anchorages with the hairpin set 50 mm below the concrete surface. The investigators concluded that this is a result of the additional bending moment experienced by the anchor bolt in the latter cas. ‘The initial rigidity of the anchorage is typically not influenced by reinforcement, However, the maximum ioad increases depending on the size and pattern of the ‘reinforcement as discussed, 5.3.3.7. Shear load parallel to the edge. If an anchor close to an edge is loaded with a shear force acting parallel to the free edge, research by Stichting Bouwresearch (1971) indicates that the failure load is about twice as large as can be expected for loading perpendicular to the edge. However, little is known about this phenomenon. 5.3.4, Failure loads — concrete crushing combined with anchor pull-out Headed anchors loaded in shear typically exhibit crushing of the concrete in front of the anchor. For anchors with a relatively small embedment depth, this may lead to a concrete fracture behind the anchor which in turn causes a pull-out failure (see Fig. 5.37(g)). While no theoretical formulations are available at present to predict the failure load associated with this type of failure, it is likely to depend principally on the concrete strength, the embedment depth, the head diameter and the spacing of the anchors, as applicable. Generally, the required embedment to preclude this type of failure is four to five times the diameter of the anchor. For high-strength steels or low-strength concrete and anchor groups, however, this embedment length may net be adequate 5.3.5. Failure loads — concrete spall ‘The primary factors influencing the shear load at which spalling of the concrete occurs are tensile strength of the concrete, flexural stiffness of the anchor shaft, anchor shaft diameter, embedment depth and deformability (E modulus) of the concrete, The corresponding shear capacity is given by Olgeard, Stutter & Fisher (1971), AISC (1978), Klingner & Mendonca (1982b) and Roik (1982) as Vim = O-5AGE) (bf, N) (5.52) Hawkins (1987) proposes the following equation based on statistical analysis of test results Vem = 18.2 (dy? f205 (1S + 1.1 Aye + dw) (Ibe) (5.53) Com 313-1 FoF BBL + LL he + dw) ON) 5.4, Combined tension and shear loading Fig. 5.57. Interaction diagram for headed anchors (est results from Bode & Hanenkamp, 1985) HEADED ANCHORS IN UNCRACKED CONCRETE where dw = washer diameter < hy. The tests were conducted on headed studs with-a washer beneath the nut. If no washer is present, dw is taken as zero in equation (5.53) ‘From tests with headed studs, the depth of the spall was found to be typically about 0.4—0.6 times the anchor diameter (Fuchs, 1990). Based on theoretical considerations, the following equation was proposed for the spalling load Vom = As(O-Uf, + 2.%fce) — (N) 6.54) Fuchs (1990) limits the applicability of equation (5.54) to anchors with an embedment depth of at least five times the anchor shaft diameter. For embedment depths of between three and five times the shaft diameter, a deduction factor is proposed Uy = O2hegld <1 (5.55) Investigations by Eligehausen & Fuchs (1988) have indicated that the ultimate oad attained by the anchor in the case of steel failure is not greatly affected by the nature of the concrete spall. 5.4.1, Steel failure "The steel failure mechanism for headed anchors in combined tension and shear is characterized by yielding and fracture of the anchor shaft due to combined shear, tension and flexural stresses. 5.4.1.1. Meihod according to ACI 349-85 (1985) (shear— friction theory). ‘The shear strength of single anchors or anchor groups with full embedment is calculated by use of shear—friction theory. This approach is valid for steel failure only. Equation (5.56) is derived frome the design equation in ACT 349-85, (1985), with a strength reduction factor @ = 1.0. N+ Vin = Ady (5.56) ihre bis ia le: pr eae | mae] ||, ‘i Lrwlo]|ofe = 22 | 100 a . ! Pet Hh 2 |io| 2 | "4 a TSS 3 N oe 2 cel BS \ Equation (5.61) Pery nee p a 0.50] A laa a \ oO 0.25) - quan (680) °o 0.25, 0.50 0.75 1.00 = 1.50 VoseeV Ves! 6 MONOTONIC LOADING IN CONCRETE 6 where 1 is the coefficient of friction (~ 0.55~—0.9, depending on the location of the anchor plate in relation to the concrete surface) (see section 5.3.2.1), A, is the stress area for all threaded anchors A, = (14) (doug = 0.9743/n)? 6.57) ris the number of threads per inch (254mm), From equation (5.56), the ACT approach to interaction gives a straight-line interaction diagram 5.4.1.2. Methods not based on shear—friction theory. The strength of steel bolts and studs subjected to tension and shear loads is commonly represented as an elliptical interaction curve. The interaction equation that defines the ultimate load thus takes the form WIN + VY" S10 6.58) cctypically varies between 5/3 (McMackin et al., 1973) and 2 (Thurner, 1985; Shaik & Whayong, 1985). 5.4.2. Concrete failure To calculate the failure load under combined tension and shear loading, three approaches are found in the literature, These may be characterized as straght- line, trilinear and elliptical functions. 5.4.2.1, Straight-line function. Based on test results for wedge-type expansion anchors, Johnson & Lew (1990) propose a straight line as a lower bound for the interaction diagram (see Fig. 5.57) (NIN,) + (WV) = 1.0 6.59) 5.4.2.2, Trilinear function. Bode & Roik (1987) propose a trilinear function for headed anchors NIN, <1 (6.604) Wy st (6.600) (WIN,) + (VIV,) 12 (5.60¢) where N and are the applied tension and shear loads respectively and Ny and V, are the ultimate tension and shear loads respectively. Figure 5.57 shows the results of tests on headed anchors in which the angle of the load was systematically varied between a = 0° (pure shear) and « = 90° (tension). The test results are presented in the form of an interaction. diagram. The measured failure loads are related to the 5% fractile of the theoretical values as calculated by use of the y method (see sections 5.2.3.3 and 5.3.3.4). ‘Some tentative experimental investigations using quadruple fastenings indicate that the interaction equations (equation (5.60)) can also be used for anchor groups (Rehm et al., 1988). However, the authors concluded that further rescarch is needed to clarify the bearing mechanism and to establish the influence of such major parameters as number and diameter of fastening elements, anchorage depth and anchor spacing on the load—-displacement behaviour and the ultimate load. In particular, they proposed tests designed to establish the transition in failure mode for anchor groups loaded in varying degrees of tension and shear. No experimental or theoretical work has yet been conducted in this area 5.4.2.3. Elliptical functions, MeMackin et al. (1973), Shaik & Whayong (1985) and Cook & Klingner (1989) propose an elliptical function to describe the strength of anchors loaded in tension and shear WINS + IV," = 1.0 6.61) where a is determined from analysis of test results. Shaik & Whayong (1985) propose a = 2, which gives a circular function HEADED ANCHORS IN-UNCRACKED CONCRETE 120) — 0} “Tonsion: KN Fig, 5.58. Tenson—shear | interaction for castin-place anchors (Cook & Klingner, °0 40 60 1989) Shear: KN (Fig. 5.57); MceMackin et al. (1973) and Cook & Klingner (1989) propose @ = 5/3, which leads to an interaction diagram in the form of a flattened ellipse (Fig. 5.57). PCI (1985) proposes the use of a = 4/3 for anchors used. ‘ in precast construction. 2 ‘Cook & Klinger (1989) measured anchor tensile forces directly for multiple- ‘ anchor connections. In the calculation of the anchor shear forces, the coefficient of friction was known and the equilibrium relations discussed in section 5.6.4 were used. As shown in Fig. 5.58, an elliptical tension—shear interaction (see also equation (5.61)) relation was fouind to provide a reasonable and generally conservative fit to the tést data. A linear tension—shear interaction relation was found to be conservative. 5.5. Bending of the To determine the ultimate strength of the steel-to-concrete connection shown baseplate in Fig. 5.59, three separate capacities must be considered (@ the flexural capacity of the attached member ' ( the flexural capacity of the steel baseplate (0) the capacity of the fastener. Fig. 5.59. Typical steel-to- a ttomont concrete connection oe Attaches membor Fastener ( n MONOTONIC LOADING IN CONCRETE 5.6. Behaviour of multiple-anchor ductile attachments to concrete If the connection is proportioned so that yielding occurs in the attached ‘member, the baseplate or the anchors before failure of the concrete, then the connection can be described as ductile. ‘The flexural capacity of the attached member can be calculated by the same means as used for structural steel design, and is not discussed here. The flexural capacity of the baseplate and the capacity of the fastener itself are discussed in this section and section 5.6 respectively. In the case of a perfectly rigid baseplate, the compression resultant is located. at the toe of the baseplate. In the case of a flexible baseplate, tests by Cook & Klingner (1989) show that the location of the compressive reaction moves closer to the tensile anchors, and is given by the condition of moment equilibrium between the applied moment and the tensile resultant of the anchors, ‘As shown in Fig. 5.60, the overhanging portion of the plate on which the compressive reaction acts is essentially a cantilever fixed at its intersection with the Compressive element of the attached member, and loaded by a movable concentrated load (the compressive reaction). The boundary conditions for this cantilever require the fixed support to rotate and displace. ‘The free end can rotate but cannot dispiace if the compressive reaction is at the toe of the plate, and can neither rotate nor displace if the compressive reaction moves in from the toe of the baseplate. ‘As shown in Fig. 5.60, the behaviour of a flexible baseplate can be described as follows. (@ Initially, the baseplate rotates as a rigid body, pivoting about the toe of the plate. () As the compressive reaction increases, the portion of the baseplate adjacent (o the compressive element of the attached member reaches the yield moment M, of the baseplate. This causes the compressive reaction C to move inwards towards the compression element. The smallest distance qi, between the compressive reaction and the ‘compression element of the attached member can be determined from Sein = MyIC 6.62) (6) With a further increase in the compressive reaction, the baseplate forms a plastic hinge M, and the compressive reaction C moves away from the compression element. The longest distance that the compressive reaction moves away from the support is determined from X= MIC (5.63) To locate the compressive reaction conservatively, it can be considered to be at a distance xj, determined by equation (5.62) from the edge of the compression element of the attached member. ‘The most extensive currently available investigation of the behaviour of ductile ‘multiple-anchor attachments to concrete is that of Cook & Klingner (1989), whose principal findings are discussed in this section. The experimental programme included the following types of test (a) coefficient of friction (©) ultimate load: two-anchor rigid baseplate, four-anchor rigid baseplate, six-anchor rigid baseplate and six-anchor flexible baseplate. 5.6.1. Coefficient of friction In the tests of Cook & Klingner (1989), the average mean value for the coefficient of friction between a surface-mounted stee! plate and hardened concrete was 0.43, with a standard deviation of 0.09. The coefficient of friction was not significantly affected by the surface condition of the concrete, the Fig. 5.60. Effect of baseplate flexibility on the location of the compressive reaction (Cook & Klingner, 1983): (a) compressive reaction at 10¢; (b) compressive reaction shifted in from toe Fig. 5.61. Typical load— displacement diagrams for ssit-anchor rigid baseplate test dominated by anchor shear (Cook & Klingner, 1989) HEADED ANCHORS IN UNCRACKED CONCRETE Meaty — of BREE MaMy je ce HL “magnitude of the compressive force or ‘digging in’ of the toc of the rigid baseplate to the concrete. For computation of capacity, the authors recommend that the coefficient of friction « be taken as 0.40, with a strength reduction factor ¢ of 0.65. . ) ® 5.6.2. Uliimate capacity of multiple-anchor attachments to concrete ‘All ultimate-load specimens of Cook & Klingner (1989) failed by yielding and fracture of the anchors. Three typical load—displacement curves (Fig. 45.61) show the applied shear as a function of the total displacement (the square root of the sum of the squares of the horizontal slip and the vertical displacement at the location of the outer row of tension anchors). Even for connections dominated by anchor shear, the anchors underwent significant inelastic deformation before failure. 5.6.3. Combined tension and shear interaction relationships ‘The descriptions given in section 5.4 are applicable. More details of multiple- anchor connections under combined tension and shear load are given in section 5.4.2.3, 5.6.4, Distribution of tension and shear among anchors Results of the multiple-anchor attachment tests of Cook & Klinger (1989) indicated that (@ tension and shear forces in the anchors redistribute inelastically as required to maintain equilibrium with the applied loading (b) for connections dominated by moment (high eccentricity of the applied load) the anchors away from the toe of the baseplate attain their full tensile strength (©) for connections dominated by shear (low eccentricity of the applied load), the ultimate strength of the connection is not sensitive to the distribution of tension in the anchors es + 4 : | es Displacement mm Binitaapuaete te MONOTONIC LOADING IN CONCRETE Table 5.2. Ranges of behaviour for ductile ‘multiple-anchor connections (Cook & Klingner, 1989): T, = the tensile strengih of the anchor @) the initial distribution of anchor tension (before inelastic redistribution) does not affect the ultimate strength of the connection. 5.6.5. Analytical basis for prediction of the strength of ductile mutiple-anchor connections According to Cook & Klingner (1989), the behaviour of a ductile multiple- anchor connection can be separated into three distinct ranges (Table 5.2). (@ Range I. if the shear strength provided by the frictional force (developed from the compressive reaction produced by the applied moment) is larger than the applied shear, anchors are not required for shear. The anchors in the tension zone can be assumed to develop their full tensile strength for moment resistance, (®) Range 2. If the shear strength provided by the frictional force and by the anchors in the compression zone exceeds the applied shear, the anchors in the tension zone can be assumed to develop their ull tensile strength for moment resistance. (© Range 3. If the shear strength provided by the frictional force and by the anchors in the compression zone is less than the applied shear, the anchors in the tension zone must transfer the remaining shear load. The strength of the anchors in the tension zone is limited by their tension—shear interaction. The transitions between these three ranges of behaviour coincide with two critical values of shear load eccentricity e, equal to the moment—shear ratio M/Y of the applied loading at the surface of the concrete, The first critical eccentricity e” represents the transition between range 1 and range 2 behaviour, and corresponds to the point at which the applied shear load is equal to the frictional force. For eccentricities larger than e’, the connection does not slip and no shear anchors are required. For eccentricities smaller than e’, the connection slips and shear anchors must be provided. ‘The second critical eccentricity e” represents the transition between range 2 and range 3 behaviour, and corresponds to the point at which the applied shear load is equal to the sum of the frictional force and the shear strength of the anchors in the compression zone. For eccentricities larger than e”, ‘the anchors in the tension zone can be assumed to develop their full tensile strength for moment resistance. For eccentricities smaller than e”, the anchors in the tension zone carry both tension and shear. If no.anchors are provided in the compression zone, range 2 behaviour is not applicable, and e" =e". Range 3 Range 2 Range 1 Connection stips Connection slips Connection does not slip Anchors in the compression zone | Anchors in the compression zone | Anchors are not requied for shear are at their maximum shear strength 77, transfer shear and can achieve their maximum shear strength ¥7, Anchors in the tension zone are in | Anchors in the tension zone can | Anchors in the tension zone can combined tension and shear and | achieve their maximum tensile achieve their maximum tensile ccan achieve their maximum strength in combined tension and shear strength T, strength T, ‘Shear load eccentricity ¢ or moment—shear ratio M/V ° 80 e e ° HEADED ANCHORS IN-UNCRACKED CONCRETE 5.6.6, Equations to predict the capacity of ductile imultiple-anchor connections The strength of connections dominated by shear depends on the tension— shear interaction of the anchors. As-derived equations are given below for elliptical and linear interaction (Cook & Klingner, 1989) For elliptical tension—shear interaction, the anchor shear strength is given by Vy = (T3 — TH)? (5.64) For the more conservative linear tension —shear interaction, the anchor shear strength is given by Va = (To — Tr) (5.65) where V, is the shear strength of an anchor in combined tension and shear and + is the ratio of the shear strength of the anchor to the tensile strength of the anchor (0.5 for cast-in-place and adhesive anchors and 0.6 for undercut anchors, see section 5.3.2.2) ‘The critical eccentricities e” and e” may be determined by the conditions of equilibrium, The following formulations for ¢’ and e” ate applicable to connections with multiple rows of anchors if d is taken as the distance from the compressive reaction C to the centroid of the anchors in the tension zone, ris taken as the number of rows of anchors in the tension zone, m is taken ‘as the number of rows of anchors in the compression zone and T, is taken as the tensile strength of a row of anchors. ‘At the minimum eccentricity e”, the applied shear load V is equal to the frictional force #C ef = dip (5.66) where e’ is the minimum eccentricity for multiple-anchor connections without shear in the anchors, iis the coefficient of friction between steel and concrete, and d is the distance from the compressive reaction to the centroid of the anchors in the tension zone. ‘At the minimum eccentricity ¢”, the applied shear load V is equal to the sum of the frictional force C and the shear strength of the rows of anchors in the compression zone myT e” = ndl(np + my) 6.6 where e” is the minimum eccentricity for multiple-anchor connections without combined tension and shear in the anchors, n is the number of rows of anchors in the tension zone, m is the number of rows of anchors in the compression zone, + i the ratio of the shear strength of the anchor to the tensile strength of the anchor, and . and d are as defined for equation (5.66). ¢” reduces to e’ when no anchors are provided in the compression zone. 5.6.6.1. Distribution of tension. Fot connections with more than one row of anchors in the tension zone, the distribution of tension cannot be adequately predicted by traditional analysis methods. Provided that sufficient anchors are present to satisfy cquilibrium, the connection will perform satisfactorily However, the issue of available inelastic deformation capacity must be addressed. The above analysis is based on the assumption that materials have infinite plastic deformation capacity after yield: this is not in fact the case. Cook & Klingner (1989) have proposed that the distance d; between the inner row of anchors and the compression reaction should not be less than ~10% of the distance d, (4, > 0.10 d, in Fig. 5.62). This is an attempt to ensure that the tensile strain in the inner row of anchors ¢ will be at least 0.01 when the tensile strain in the outer row of anchors ¢ reaches its maximum value ¢,. Anchor materials typically have a specified minimum elongation requirement of at least 10% in 50mm. This represents an ultimate a

You might also like